This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Scaling Robust Optimization for Multi-Agent Robotic Systems: A Distributed Perspective

Arshiya Taj Abdul1, Augustinos D. Saravanos1 and Evangelos A. Theodorou Georgia Institute of Technology, GA, USA {aabdul6, asaravanos, evangelos.theodorou}@gatech.edu 1These authors contributed equally
Abstract

This paper presents a novel distributed robust optimization scheme for steering distributions of multi-agent systems under stochastic and deterministic uncertainty. Robust optimization is a subfield of optimization which aims in discovering an optimal solution that remains robustly feasible for all possible realizations of the problem parameters within a given uncertainty set. Such approaches would naturally constitute an ideal candidate for multi-robot control, where in addition to stochastic noise, there might be exogenous deterministic disturbances. Nevertheless, as these methods are usually associated with significantly high computational demands, their application to multi-agent robotics has remained limited. The scope of this work is to propose a scalable robust optimization framework that effectively addresses both types of uncertainties, while retaining computational efficiency and scalability. In this direction, we provide tractable approximations for robust constraints that are relevant in multi-robot settings. Subsequently, we demonstrate how computations can be distributed through an Alternating Direction Method of Multipliers (ADMM) approach towards achieving scalability and communication efficiency. Simulation results highlight the performance of the proposed algorithm in effectively handling both stochastic and deterministic uncertainty in multi-robot systems. The scalability of the method is also emphasized by showcasing tasks with up to 100 agents. The results of this work indicate the promise of blending robust optimization, distribution steering and distributed optimization towards achieving scalable, safe and robust multi-robot control.

I Introduction

Multi-agent optimization and control problems are emerging increasingly often in a vast variety of robotics applications such as multi-vehicle coordination [16, 43], swarm robotics [37], cooperative manipulation [13] and others. As the size and complexity of such systems has been rapidly escalating, a substantial requirement for developing scalable and distributed algorithms for tackling these problems has been developing as well [35, 38, 46]. In the meantime, accompanying such algorithms with safety and robustness guarantees in environments with substantial levels of uncertainty is also a highly desirable attribute for securing the safe operation of all agents.

Uncertainty has been primarily modeled as stochastic noise in a significant portion of the multi-agent control literature. Some indicative approaches involve multi-agent LQG control [1, 24, 39], distributed covariance steering [33, 34], multi-agent path integral control [14, 42] and learning-based stochastic optimal control methods [21, 26], among others. Nevertheless, such approaches would typically fall short in scenarios where the underlying disturbances might be exogenous interventions, environmental factors, modelling inaccuracies, etc., that cannot be properly represented as stochastic signals [5]. Such problems create the necessity of finding robust optimal policies that would guarantee feasibility for all possible values of such disturbances within some bounds.

Robust optimization is a subfield of optimization that addresses the problem of discovering an optimal solution that remains robust feasible for all possible realizations of the problem parameters within a given uncertainty set [5, 6]. Although robust optimization has found significant successful applications in many areas such as power systems [8], supply chain networks [29], communication systems [10], etc., there have only been limited attempts for addressing problems relevant to robotics [19, 23, 30, 40]. The field of robust control is a also closely related area with long history in addressing disturbances from predefined sets [12, 45]. However, as this area is mainly focused on synthesis techniques for stabilization and performance maintenance, merging such methods with distributed optimization and steering state distributions for multi-agent systems is not as straightforward. On the contrary, as robust optimization allows for more flexibility in designing such approaches, the current work is focused on leveraging the latter for safe multi-agent control.

Another promising area that has recently emerged in the context of safe stochastic control is the one of steering the state distributions of systems under predefined target distributions - mostly known as covariance steering [3, 11, 22]. Such methods have found several successful applications in path planning [25], trajectory optimization [4, 44], robotic manipulation [20], flight control [7, 31], multi-robot control [33, 34] and others. More recently, a convex optimization framework for steering state distributions and effectively addressing both stochastic noise and bounded deterministic disturbances was presented [19]. In a similar direction, a data-driven robust optimization-based method was recently proposed in [28], including estimating the noise realization from collected data. Nevertheless, as such approaches result in high-dimensional semidefinite programming (SDP) problems, the computational demands for addressing multi-agent control tasks have remained prohibitive.

The scope of this work is to provide a scalable robust optimization framework for steering distributions of multi-agent systems to specific targets under stochastic and deterministic uncertainty. Towards this direction, we present a distributed algorithm for safe and robust multi-robot control whose effectiveness and scalability are verified through simulation experiments. The specific contributions of this paper can be listed as follows.

  • We combine elements from the areas of robust optimization, distribution steering and distributed computation architectures for deriving a scalable and robust approach for multi-agent control. To our best knowledge, this is the first work to fuse these areas into a single methodology.

  • We circumvent computational expensiveness by providing tractable approximations of robust optimization constraints that are relevant for multi-robot control problems.

  • We present a distributed algorithm based on an Alternating Direction Method of Multipliers (ADMM) approach for multi-agent robust distribution steering.

  • We provide simulation results that demonstrate the effectiveness of the proposed algorithm in handling stochastic and deterministic uncertainty. The scalability of the method is also highlighted by showing multi-robot tasks with up to 100 agents.

The remaining of this paper is organized as follows. In Section II, we state the multi-agent problem to be considered in this work. The proposed distributed robust optimization framework is presented in Section III. An analysis of performance and scalability is demonstrated through simulation experiments in Section IV. Finally, the conclusion of this paper, as well as future research directions, are provided in Section V.

II Problem Statement

In this section, we present the multi-agent robust optimization problem to be addressed in this work. Initially, we familiarize the reader with the notation used throughout this paper. Subsequently, we present the basic multi-agent problem setup and the dynamics of the agents. The two main characterizations of underlying uncertainty, i.e., deterministic and stochastic, are then formulated. Then, we specify the control policies of the agents and the robust constraints that are considered in our framework. Finally, the exact formulation of the multi-agent problem is provided.

II-A Notation

The integer set [a,b][a,b]\cap\mathbb{Z} is denoted with a,b\llbracket a,b\rrbracket. The space of matrices 𝐗n×n{\bf X}\in\mathbb{R}^{n\times n} that are symmetric positive (semi)-definite, i.e., 𝐗0{\bf X}\succ 0 (𝐗0{\bf X}\succeq 0) is denoted with 𝕊n++\mathbb{S}_{n}^{++} (𝕊n+\mathbb{S}_{n}^{+}). The inner product of two vectors 𝒙,𝒚n{\bm{x}},{\bm{y}}\in\mathbb{R}^{n} is denoted with 𝒙,𝒚=𝒙T𝒚\langle{\bm{x}},{\bm{y}}\rangle={\bm{x}}^{\mathrm{T}}{\bm{y}}. The 2\ell_{2}-norm of a vector 𝒙=[x1xn]n,{\bm{x}}=\begin{bmatrix}x_{1}&\dots&x_{n}\end{bmatrix}\in\mathbb{R}^{n}, is defined as 𝒙2=𝒙,𝒙\|{\bm{x}}\|_{2}=\sqrt{\langle{\bm{x}},{\bm{x}}\rangle}. The weighted norm 𝒙𝐐=𝑸1/2𝒙2=𝒙T𝑸𝒙\|{\bm{x}}\|_{{\bf Q}}=\|{\bm{Q}}^{1/2}{\bm{x}}\|_{2}=\sqrt{{\bm{x}}^{\mathrm{T}}{\bm{Q}}{\bm{x}}} is also defined for any 𝐐𝕊n++{\bf Q}\in\mathbb{S}_{n}^{++}. In addition, the Frobenius norm of a matrix 𝐗m×n{\bf X}\in\mathbb{R}^{m\times n} is given by 𝐗F=tr(𝐗T𝐗)\|{\bf X}\|_{F}=\sqrt{\mathrm{tr}({\bf X}^{\mathrm{T}}{\bf X})}. With [𝒙1;;𝒙n][{\bm{x}}_{1};\dots;{\bm{x}}_{n}], we denote the vertical concatenation of a series of vectors 𝒙1,,𝒙n{\bm{x}}_{1},\dots,{\bm{x}}_{n}. Furthermore, the probability of an event A\pazocal{A} is denoted with (A)\mathbb{P}(\pazocal{A}). Given a random variable (r.v.) 𝒙n{\bm{x}}\in\mathbb{R}^{n}, its expectation and covariance are denoted with 𝔼[x]n\mathbb{E}[x]\in\mathbb{R}^{n} and Cov(𝒙)𝕊n+\mathrm{Cov}({\bm{x}})\in\mathbb{S}_{n}^{+}, respectively. The cardinality of a set X\pazocal{X} is denoted as n(X)n(\pazocal{X}). Finally, given a set X\pazocal{X}, the indicator function IX\pazocal{I}_{\pazocal{X}} is defined as IX(x)=0\pazocal{I}_{\pazocal{X}}(x)=0, if xXx\in\pazocal{X}, and IX(x)=+\pazocal{I}_{\pazocal{X}}(x)=+\infty, otherwise.

II-B Problem Setup

Let us consider a multi-agent system of NN agents defined by the set V={1,,N}\pazocal{V}=\{1,\dots,N\}. All agents iVi\in\pazocal{V} may be subject to diverse dynamics, inter-agent interactions, and exogenous deterministic and stochastic uncertainty. Each agent iVi\in\pazocal{V} has a set of neighbors defined by NiV\pazocal{N}_{i}\subseteq\pazocal{V}. Furthermore, for each agent iVi\in\pazocal{V}, we define the set of agents that consider ii as a neighbor through Pi={jV|iNj}\pazocal{P}_{i}=\{j\in\pazocal{V}|i\in\pazocal{N}_{j}\}.

II-C Dynamics

We consider the following discrete-time linear time-varying dynamics for each agent iVi\in\pazocal{V},

xk+1i=Akixki+Bkiuki+Ckidki+Dkiwki,k0,T1,x_{k+1}^{i}=A_{k}^{i}x_{k}^{i}+B_{k}^{i}u_{k}^{i}+C_{k}^{i}d_{k}^{i}+D_{k}^{i}w_{k}^{i},~{}k\in\llbracket 0,T-1\rrbracket, (1)

where xkinxix_{k}^{i}\in\mathbb{R}^{n_{x_{i}}} is the state, ukinuiu_{k}^{i}\in\mathbb{R}^{n_{u_{i}}} is the control input of agent ii at time kk, and TT is the time horizon. The terms dkindid_{k}^{i}\in\mathbb{R}^{n_{d_{i}}} represent exogenous deterministic disturbances, while the terms wiknwiw_{i}^{k}\in\mathbb{R}^{n_{w_{i}}} refer to stochastic disturbances, with their exact mathematical representations presented in Section II-D. Each initial state x0i,iVx_{0}^{i},~{}i\in\pazocal{V}, is of the following form

x0i=x¯0i+d¯0i+s0i,x_{0}^{i}=\bar{x}_{0}^{i}+\bar{d}_{0}^{i}+s_{0}^{i}, (2)

where x¯0i\bar{x}_{0}^{i} is the known part, while d¯0i\bar{d}_{0}^{i} and s0is_{0}^{i} refer to the unknown exogenous deterministic and stochastic parts, respectively. Finally, Aki,Bki,Cki,DkiA_{k}^{i},B_{k}^{i},C_{k}^{i},D_{k}^{i} are the dynamics matrices of appropriate dimensions.

For convenience, let us also define the sequences 𝒙i=[x0i;;xTi](T+1)nxi{\bm{x}}^{i}=[x_{0}^{i};\dots;x_{T}^{i}]\in\mathbb{R}^{(T+1)n_{x_{i}}}, 𝒖i=[u0i;;uT1i]{\bm{u}}^{i}=[u_{0}^{i};\dots;u_{T-1}^{i}] Tnui\in\mathbb{R}^{Tn_{u_{i}}}, 𝒘i=[s0i;w0i;;wT1i]nxi+Tnwi{\bm{w}}^{i}=[s_{0}^{i};w_{0}^{i};\dots;w_{T-1}^{i}]\in\mathbb{R}^{n_{x_{i}}+Tn_{w_{i}}} and 𝜻i=[d¯0i;d0i;;dT1i]nxi+Tndi\bm{\zeta}^{i}=[\bar{d}_{0}^{i};d_{0}^{i};\dots;d_{T-1}^{i}]\in\mathbb{R}^{n_{x_{i}}+Tn_{d_{i}}}. The dynamics (1) can then be rewritten in a more compact form as

𝒙i=𝐆0ix¯0i+𝐆ui𝒖i+𝐆𝒘i𝒘i+𝐆ζi𝜻i,{\bm{x}}^{i}={\bf G}_{0}^{i}\bar{x}_{0}^{i}+{\bf G}_{u}^{i}{\bm{u}}^{i}+{\bf G}_{{\bm{w}}}^{i}\bm{{\bm{w}}}^{i}+{\bf G}_{\zeta}^{i}\bm{\zeta}^{i}, (3)

where the matrices 𝐆0i,𝐆ui,𝐆𝒘i,𝐆ζi{\bf G}_{0}^{i},{\bf G}_{u}^{i},{\bf G}_{{\bm{w}}}^{i},{\bf G}_{\zeta}^{i} are provided in Section I of the Supplementary Material (SM).

II-D Characterizations of Uncertainty

As previously mentioned, we consider the existence of both exogenous deterministic and stochastic uncertainties. These two types of uncertainty are formally described as follows.

Exogenous deterministic uncertainty

The unknown deterministic disturbances 𝜻i\bm{\zeta}^{i} typically stem from external factors that cannot be accurately represented through stochastic signals. This creates a requirement to formulate optimization problems and achieve optimal policies that are feasible in a robust sense for all possible values of such disturbances in a bounded uncertainty set. Examples of such uncertainty sets include ellipsoids, polytopes, and others [5]. In this work, we assume the sequence 𝜻i\bm{\zeta}^{i} to be lying inside an ellipsoidal uncertainty set defined as follows

Ui[τi]={𝜻inζi\displaystyle\pazocal{U}_{i}[\tau^{i}]=\big{\{}\bm{\zeta}^{i}\in\mathbb{R}^{n_{\zeta_{i}}} |(𝒛in¯i,τi):\displaystyle|\;\exists({\bm{z}}_{i}\in\mathbb{R}^{\bar{n}_{i}},\tau^{i}\in\mathbb{R}):\; (4)
𝜻i=𝚪i𝒛i,𝐒i𝒛i,𝒛iτi},\displaystyle\bm{\zeta}^{i}=\bm{\Gamma}_{i}{\bm{z}}_{i},\;\langle{\bf S}_{i}{\bm{z}}_{i},{\bm{z}}_{i}\rangle\leq\tau^{i}\big{\}},

where nζi=nxi+Tndin_{\zeta_{i}}=n_{x_{i}}+Tn_{d_{i}}, 𝚪inζi×n¯i\bm{\Gamma}_{i}\in\mathbb{R}^{n_{\zeta_{i}}\times\bar{n}_{i}}, 𝐒i𝕊++n¯i{\bf S}_{i}\in\mathbb{S}_{++}^{\bar{n}_{i}} and τi>0\tau^{i}>0 is the uncertainty level. Ellipsoidal sets are used to model uncertainty in various robust control applications [27]. Further, ellipsoid sets have simple geometric structure, and more complex sets can be approximated using ellipsoids.

Stochastic uncertainty

The unknown stochastic component 𝒘i{\bm{w}}^{i} is assumed to be a Gaussian random vector with mean 𝝁𝒘i=𝔼[𝒘i]=0{\bm{\mu}}_{{{\bm{w}}}^{i}}=\mathbb{E}[{\bm{w}}^{i}]=0, and covariance 𝚺𝒘i=Cov[𝒘i]{\bm{\Sigma}}_{\bm{{\bm{w}}}^{i}}=\mathrm{Cov}[{\bm{w}}^{i}].

II-E Control Policies

Towards addressing both types of uncertainty, we consider affine purified state feedback policies [5, 19] for all agents. Let us first introduce the disturbance-free states x^ki\hat{x}_{k}^{i}, whose dynamics are given by

x^k+1i\displaystyle\hat{x}_{k+1}^{i} =Akix^ti+Bkiuki,k0,T1,\displaystyle=A_{k}^{i}\hat{x}_{t}^{i}+B_{k}^{i}u_{k}^{i},\quad k\in\llbracket 0,T-1\rrbracket, (5)
x^0i\displaystyle\hat{x}_{0}^{i} =x¯0i.\displaystyle=\bar{x}_{0}^{i}. (6)

Subsequently, we can define the purified states δki=xkix^ki\delta_{k}^{i}=x_{k}^{i}-\hat{x}_{k}^{i}, whose dynamics are given by

δk+1i\displaystyle\delta_{k+1}^{i} =Akiδki+Ckidki+Dkiwkik0,T1,\displaystyle=A_{k}^{i}\delta_{k}^{i}+C_{k}^{i}d_{k}^{i}+D_{k}^{i}w_{k}^{i}\quad k\in\llbracket 0,T-1\rrbracket, (7)
δ0i\displaystyle\delta_{0}^{i} =d¯0i+s0i,\displaystyle=\bar{d}_{0}^{i}+s_{0}^{i},

or more compactly by

𝜹i=𝐆𝒘i𝒘i+𝐆ζi𝜻i,{\bm{\delta}}^{i}={\bf G}_{{\bm{w}}}^{i}\bm{{\bm{w}}}^{i}+{\bf G}_{\zeta}^{i}\bm{\zeta}^{i}, (8)

where 𝜹i=[𝜹0i;;𝜹Ti]{\bm{\delta}}^{i}=\begin{bmatrix}{\bm{\delta}}_{0}^{i};&\dots;&{\bm{\delta}}_{T}^{i}\end{bmatrix}. We consider the following affine purified state feedback control policy for each agent ii,

uki=u¯ki+=0kKk,iδi,k0,T1,u^{i}_{k}=\bar{u}^{i}_{k}+\sum_{\ell=0}^{k}K^{i}_{k,\ell}\delta^{i}_{\ell},\;\forall\;k\in\llbracket 0,T-1\rrbracket, (9)

where u¯kinui\bar{u}^{i}_{k}\in\mathbb{R}^{n_{u_{i}}} are the feed-forward control inputs and Kk,inui×nxiK^{i}_{k,\ell}\in\mathbb{R}^{n_{u_{i}}\times n_{x_{i}}} are feedback gains on the purified states. The above policies can be written in compact form as

𝒖i=𝒖¯i+𝐊i𝜹i,{\bm{u}}^{i}=\bar{{\bm{u}}}^{i}+{\bf K}^{i}{\bm{\delta}}^{i}, (10)

where

𝒖¯i=[u¯0i;;u¯T1i],\displaystyle\bar{{\bm{u}}}^{i}=\begin{bmatrix}\bar{u}_{0}^{i};&\dots;&\bar{u}_{T-1}^{i}\end{bmatrix},
𝐊i=[K0,0i𝟎𝟎𝟎K1,0iK1,1i𝟎𝟎KT1,0iKT1,1iKT1,T1i𝟎].\displaystyle{\bf K}^{i}=\begin{bmatrix}K_{0,0}^{i}&{\bf 0}&\dots&{\bf 0}&{\bf 0}\\ K_{1,0}^{i}&K_{1,1}^{i}&\dots&{\bf 0}&{\bf 0}\\ \vdots&\vdots&\vdots&\vdots&\vdots\\ K_{T-1,0}^{i}&K_{T-1,1}^{i}&\dots&K_{T-1,T-1}^{i}&{\bf 0}\end{bmatrix}.

II-F Robust Constraints

Next, we introduce the types of constraints that are considered in this work. We highlight that the following constraints must be satisfied robustly for all possible values of 𝜻i{\bm{\zeta}}^{i} lying inside the uncertainty sets Ui,iV\pazocal{U}_{i},\ i\in\pazocal{V}.

Robust linear mean constraints

The first class of constraints we consider are robust linear constraints on the expectation of the state sequence 𝒙i{\bm{x}}^{i} of agent iVi\in\pazocal{V},

𝐀i𝔼[𝒙i]𝒃i,𝜻iUi,{\bf A}_{i}\mathbb{E}[{\bm{x}}^{i}]\leq\bm{b}_{i},\quad\forall{\bm{\zeta}}^{i}\in\pazocal{U}_{i}, (11)

where 𝐀im×nxi{\bf A}_{i}\in\mathbb{R}^{m\times n_{x_{i}}}, 𝒃im\bm{b}_{i}\in\mathbb{R}^{m}. In a robotics setting, such constraints could typically represent bounds on the mean position, velocity, etc.

Robust nonconvex norm-of-mean constraints

Furthermore, we consider nonconvex inequality constraints of the form

𝐀i𝔼[xki]𝒃i2ci,𝜻iUi,\|{\bf A}_{i}\mathbb{E}[x_{k}^{i}]-\bm{b}_{i}\|_{2}\geq c_{i},\quad\forall{\bm{\zeta}}^{i}\in\pazocal{U}_{i}, (12)

where 𝐀im×nxi{\bf A}_{i}\in\mathbb{R}^{m\times n_{x_{i}}}, 𝒃im\bm{b}_{i}\in\mathbb{R}^{m}, ci+c_{i}\in\mathbb{R}^{+}. Such constraints are typically useful for capturing avoiding obstacles, forbidden regions, etc. In addition, we extend to the inter-agent version of these constraints which is given by

𝐀i𝔼[𝒙ki]𝐀j𝔼[𝒙kj]2cij,𝜻iUi,𝜻jUj,jNi\|{\bf A}_{i}\mathbb{E}[{\bm{x}}^{i}_{k}]-{\bf A}_{j}\mathbb{E}[{\bm{x}}^{j}_{k}]\|_{2}\geq c_{ij},\quad\forall{\bm{\zeta}}^{i}\in\pazocal{U}_{i},~{}{\bm{\zeta}}^{j}\in\pazocal{U}_{j},~{}j\in\pazocal{N}_{i} (13)

where 𝐀im×nxi{\bf A}_{i}\in\mathbb{R}^{m\times n_{x_{i}}}, 𝐀jm×nxj{\bf A}_{j}\in\mathbb{R}^{m\times n_{x_{j}}}, cij+c_{ij}\in\mathbb{R}^{+}. This class of constraints typically represents collision avoidance constraints between different robots.

Robust linear chance constraints

Subsequently, we consider robust chance constraints on the states of the agents,

(𝒂iT𝒙i>bi)p,𝜻iUi,\mathbb{P}(\ {\bm{a}}_{i}^{\mathrm{T}}{\bm{x}}^{i}>b_{i})\leq p,\quad\forall{\bm{\zeta}}^{i}\in\pazocal{U}_{i}, (14)

where pp is the probability threshold. It should be noted that these constraints are also affected by the stochastic noise since they probabilistic constraints on the entire state and not only on the state mean.

Robust covariance constraints

Additionally, we consider the following constraints on the covariance matrices of the state sequences,

Cov(𝒙i)𝚺i,𝜻iUi,\mathrm{Cov}({\bm{x}}^{i})\preceq{\bm{\Sigma}}^{i},\quad\forall{\bm{\zeta}}^{i}\in\pazocal{U}_{i}, (15)

where 𝚺i{\bm{\Sigma}}^{i} is a prespecified allowed upper bound on the state covariance.

II-G Problem Formulation

Finally, we present the mathematical problem formulation for the multi-agent robust optimization problem under both stochastic and bounded deterministic uncertainties.

Problem 1 (Multi-Agent Robust Optimization Problem).

For all agents iVi\in\pazocal{V}, find the robust optimal policies 𝐮¯i,𝐊i\bar{{\bm{u}}}^{i},{\bf K}^{i}, such that

{𝒖¯i,𝐊i}iV=argminiVJi(𝒖¯i,𝐊i)\displaystyle~{}~{}~{}~{}~{}~{}~{}~{}\{\bar{{\bm{u}}}^{i},{\bf K}^{i}\}_{i\in\pazocal{V}}=\operatornamewithlimits{argmin}\sum_{i\in\pazocal{V}}J_{i}(\bar{{\bm{u}}}^{i},{\bf K}^{i})
s.t.\displaystyle\mathrm{s.t.}\; xk+1i=Akixki+Bkiuki+Ckidki+Dkiwki,k0,T1,\displaystyle~{}x_{k+1}^{i}=A_{k}^{i}x_{k}^{i}+B_{k}^{i}u_{k}^{i}+C_{k}^{i}d_{k}^{i}+D_{k}^{i}w_{k}^{i},~{}k\in\llbracket 0,T-1\rrbracket,
x¯0i : given,\displaystyle~{}\bar{x}_{0}^{i}\text{ : given}, (16)
(11)(15)\displaystyle\eqref{Robust Linear Constraints}-\eqref{Covariance Constraints}

where each cost Ji(𝐮¯i,𝐊i)=𝐮¯i𝐑u¯T𝐮¯i+𝐑K𝐊iF2J_{i}(\bar{{\bm{u}}}^{i},{\bf K}^{i})=\bar{{\bm{u}}}^{i}{}^{\mathrm{T}}{\bf R}_{\bar{u}}\bar{{\bm{u}}}^{i}+||{\bf R}_{K}{\bf K}^{i}||_{F}^{2} penalizes the control effort of agent iVi\in\pazocal{V}.

III Distributed Robust Optimization under Deterministic and Stochastic Uncertainty

In this section, we propose a distributed robust optimization framework for solving the multi-agent problem presented in Section II. Initially, we provide computationally tractable approximations for the robust constraints in Section III-A. Subsequently, Section III-B presents the new transformed problem to be solved based on the constraint approximations. Finally, in Section III-C, we present a distributed algorithm for solving the transformed problem in a scalable manner.

III-A Reformulation of Semi-infinite Constraints

We start by analyzing the effect of the disturbances on the system dynamics. In particular, since the robust constraints (11) - (13) depend on 𝔼[𝒙i]\mathbb{E}[{\bm{x}}^{i}], let us first examine the expression for 𝔼[𝒙i]\mathbb{E}[{\bm{x}}^{i}] given by

𝔼[𝒙i]=𝐆0i𝒙¯0i+𝐆ui𝒖¯i+(𝐆ui𝐊i+𝐈)𝐆ζi𝜻i,\displaystyle\mathbb{E}[{\bm{x}}^{i}]={\bf G}_{0}^{i}\bar{{\bm{x}}}_{0}^{i}+{\bf G}_{u}^{i}\bar{{\bm{u}}}^{i}+({\bf G}_{u}^{i}{\bf K}^{i}+{\bf I}){\bf G}_{\zeta}^{i}\bm{\zeta}^{i}, (17)

since 𝔼[𝒘i]=0\mathbb{E}[{\bm{w}}^{i}]=0 (derivation can be referenced from Section II of the SM). It is important to emphasize that the mean state relies only on the deterministic - and not the stochastic - disturbances. For convenience, we also define

𝝁x,u¯i=𝐆0i𝒙¯0i+𝐆ui𝒖¯i,𝐌i=(𝐆ui𝐊i+𝐈)𝐆ζi,{\bm{\mu}}_{x,\bar{u}}^{i}={\bf G}_{0}^{i}\bar{{\bm{x}}}_{0}^{i}+{\bf G}_{u}^{i}\bar{{\bm{u}}}^{i},\quad{\bf M}_{i}=({\bf G}_{u}^{i}{\bf K}^{i}+{\bf I}){\bf G}_{\zeta}^{i}, (18)

as well as the matrix 𝐏ki{\bf P}_{k}^{i} such that xki=𝐏ki𝒙ix_{k}^{i}={\bf P}_{k}^{i}{\bm{x}}^{i}, 𝐌~ki=𝐏ki𝐌i\tilde{{\bf M}}_{k}^{i}={\bf P}^{i}_{k}{\bf M}_{i}, and 𝝁xk,u¯i=𝐏ki𝝁x,u¯i{\bm{\mu}}_{x_{k},\bar{u}}^{i}={\bf P}^{i}_{k}{\bm{\mu}}_{x,\bar{u}}^{i}. Subsequently, we proceed with providing tractable approximations for the robust constraints presented in Section II.

III-A1 Robust linear mean constraints

Let us consider a single constraint from the set of linear constraints (11), given by

𝒂iT𝔼[𝒙i]bi,𝜻iUi.\displaystyle{\bm{a}}_{i}^{\mathrm{T}}\mathbb{E}[{\bm{x}}^{i}]\leq b_{i},\quad\forall{\bm{\zeta}}^{i}\in\pazocal{U}_{i}. (19)

Note that the above constraint can be reformulated as

max𝜻iUi𝒂iT𝔼[𝒙i]bi.\displaystyle\max_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i}}{\bm{a}}_{i}^{\mathrm{T}}\mathbb{E}[{\bm{x}}^{i}]\leq b_{i}. (20)

The following proposition provides the exact optimal value of the maximization problem in the LHS of (20).

Proposition 1.

The maximum value of 𝐚iT𝔼[𝐱i]{\bm{a}}_{i}^{\mathrm{T}}\mathbb{E}[{\bm{x}}^{i}] when the uncertainty vector 𝛇i{\bm{\zeta}}^{i} lies in the set Ui\pazocal{U}_{i} is given by

max𝜻iUi𝒂iT𝔼[𝒙i]=𝒂iT𝝁x,u¯i+(τi)1/2𝚪iT𝐌iT𝒂i𝐒i1.\max_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i}}{\bm{a}}_{i}^{\mathrm{T}}\mathbb{E}[{\bm{x}}^{i}]={\bm{a}}_{i}^{\mathrm{T}}{\bm{\mu}}_{x,\bar{u}}^{i}+(\tau^{i})^{1/2}||\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}{\bm{a}}_{i}||_{{\bf S}_{i}^{-1}}. (21)
Proof.

Provided in Section II of the SM. ∎

Using Proposition 1, an equivalent tractable version of the constraint (20) can be given as follows

𝒂iT𝝁x,u¯i+(τi)1/2𝚪iT𝐌iT𝒂i𝐒i1bi.{\bm{a}}_{i}^{\mathrm{T}}{\bm{\mu}}_{x,\bar{u}}^{i}+(\tau^{i})^{1/2}||\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}{\bm{a}}_{i}||_{{\bf S}_{i}^{-1}}\leq b_{i}. (22)

Similar to the maximum value case, the minimum value can be provided as

min𝜻iUi𝒂iT𝔼[𝒙i]=𝒂iT𝝁x,u¯i(τi)1/2𝚪iT𝐌iT𝒂i𝐒i1.\displaystyle\min_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i}}{\bm{a}}_{i}^{\mathrm{T}}\mathbb{E}[{\bm{x}}^{i}]={\bm{a}}_{i}^{\mathrm{T}}{\bm{\mu}}_{x,\bar{u}}^{i}-(\tau^{i})^{1/2}||\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}{\bm{a}}_{i}||_{{\bf S}_{i}^{-1}}. (23)

Based on the above observations, we introduce the following additional form of robust linear mean constraints

𝒂¯iT𝝁x,u¯i=b¯i,(τi)1/2𝚪iT𝐌iT𝒂¯i𝐒i1ϵi,\displaystyle\bar{{\bm{a}}}_{i}^{\mathrm{T}}{\bm{\mu}}_{x,\bar{u}}^{i}=\bar{b}_{i},\quad(\tau^{i})^{1/2}||\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}\bar{{\bm{a}}}_{i}||_{{\bf S}_{i}^{-1}}\leq\epsilon_{i}, (24)

with 𝒂¯inxi,b¯i\bar{{\bm{a}}}_{i}\in\mathbb{R}^{n_{x_{i}}},\bar{b}_{i}\in\mathbb{R} and ϵi+\epsilon_{i}\in\mathbb{R}^{+}, which would imply the following constraints

𝒂¯iT𝔼[𝒙i]b¯i+ϵi,𝒂¯iT𝔼[𝒙i]b¯iϵi,𝜻iUi.\displaystyle\bar{{\bm{a}}}_{i}^{\mathrm{T}}\mathbb{E}[{\bm{x}}^{i}]\leq\bar{b}_{i}+\epsilon_{i},\quad\bar{{\bm{a}}}_{i}^{\mathrm{T}}\mathbb{E}[{\bm{x}}^{i}]\geq\bar{b}_{i}-\epsilon_{i},\quad\forall{\bm{\zeta}}^{i}\in\pazocal{U}_{i}. (25)

The constraint form (24) allows us to decouple the mean constraint in terms of 𝒖¯i\bar{{\bm{u}}}^{i} and 𝐊i{\bf K}^{i}, such that the expectation of the disturbance-free state 𝝁x,u¯i{\bm{\mu}}_{x,\bar{u}}^{i} can be controlled using the feed-forward control 𝒖¯i\bar{{\bm{u}}}^{i}, and the expectation of the state mean component 𝐌i𝜻i{\bf M}_{i}{\bm{\zeta}}^{i} dependent on the uncertainty can be controlled using the feedback control gain 𝐊i{\bf K}^{i}.

III-A2 Robust nonconvex norm-of-mean constraints

Subsequently, we consider the robust nonlinear constraint (12). We present two approaches for deriving the equivalent/approximate tractable constraints to the above constraint.

Initial Approach

The robust constraint in (12) can be rewritten as

min𝜻iUi𝐀i𝔼[xki]𝒃i2ci.\displaystyle\min_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i}}\|{\bf A}_{i}\mathbb{E}[x_{k}^{i}]-\bm{b}_{i}\|_{2}\geq c_{i}. (26)

Although the problem in the LHS is convex, the equivalent tractable constraints to the above, obtained by applying S-lemma [19], might not be convex due to the non-convexity of the overall constraint (details can be referenced from Section III of the SM). Thus, we first linearize the constraint (12) with respect to control parameters 𝒖¯i,𝐊i\bar{{\bm{u}}}^{i},{\bf K}^{i} around the nominal values 𝒖¯i,l,𝐊i,l\bar{{\bm{u}}}^{i,l},{\bf K}^{i,l}. Through this approach, we obtain the following constraint

𝜻i𝒬T(𝐊i,𝐊i,l)𝜻i+2\displaystyle\bm{\zeta}^{i}{}^{\mathrm{T}}\mathscr{Q}({\bf K}^{i},{\bf K}^{i,l})\bm{\zeta}^{i}+2 𝒬¯(𝐊i,𝐊i,l,𝒖¯i,𝒖¯i,l)𝜻i\displaystyle\bar{\mathscr{Q}}({\bf K}^{i},{\bf K}^{i,l},\bar{{\bm{u}}}^{i},\bar{{\bm{u}}}^{i,l})\bm{\zeta}^{i} (27)
+q(𝒖¯i,𝒖¯i,l)ci2,𝜻iUi,\displaystyle+q(\bar{{\bm{u}}}^{i},\bar{{\bm{u}}}^{i,l})\geq c_{i}^{2},\quad\forall\;\bm{\zeta}^{i}\in\pazocal{U}_{i},

where 𝒬(𝐊i,𝐊i,l),𝒬¯(𝐊i,𝐊i,l,𝒖¯i,𝒖¯i,l)\mathscr{Q}({\bf K}^{i},{\bf K}^{i,l}),\bar{\mathscr{Q}}({\bf K}^{i},{\bf K}^{i,l},\bar{{\bm{u}}}^{i},\bar{{\bm{u}}}^{i,l}) and q(𝒖¯i,𝒖¯i,l)q(\bar{{\bm{u}}}^{i},\bar{{\bm{u}}}^{i,l}) are provided in Section III of the SM. The equivalent tractable constraints to the above constraint can be given as [19] -

β0,\displaystyle\beta\geq 0, (28)
[𝚪iT𝒬(𝐊i,𝐊i,l)𝚪i+β𝐒i𝚪iT𝒬¯(𝐊i,𝐊i,l,𝒖¯i,𝒖¯i,l)T𝒬¯(𝐊i,𝐊i,l,𝒖¯i,𝒖¯i,l)𝚪iq(𝒖¯i,𝒖¯i,l)ci2βτi]0.\displaystyle\begin{bmatrix}\bm{\Gamma}_{i}^{\mathrm{T}}\mathscr{Q}({\bf K}^{i},{\bf K}^{i,l})\bm{\Gamma}_{i}+\beta{\bf S}_{i}&\bm{\Gamma}_{i}^{\mathrm{T}}\bar{\mathscr{Q}}({\bf K}^{i},{\bf K}^{i,l},\bar{{\bm{u}}}^{i},\bar{{\bm{u}}}^{i,l})^{\mathrm{T}}\\ \bar{\mathscr{Q}}({\bf K}^{i},{\bf K}^{i,l},\bar{{\bm{u}}}^{i},\bar{{\bm{u}}}^{i,l})\bm{\Gamma}_{i}&q(\bar{{\bm{u}}}^{i},\bar{{\bm{u}}}^{i,l})-c_{i}^{2}-\beta\tau^{i}\end{bmatrix}\succeq 0.

The derived equivalent tractable constraints can be used in cases where the number of robust non-convex constraints of the defined form is relatively small. However, in applications that involve a significant number of such constraints (e.g., collision avoidance constraints), this formulation might prove to be significantly expensive computationally. For instance, consider an obstacle avoidance scenario, where each agent ii needs to enforce the constraints of the above form for each time step. Then, each agent ii in an environment with nobsn_{obs} obstacles would have nobsTn_{obs}T number of constraints of the form (28). Each constraint of the form (28) involves a matrix of size n¯i+1\bar{n}_{i}+1 (when 𝚪i=I\bm{\Gamma}_{i}=I, n¯i=nxi+Tndi\bar{n}_{i}=n_{x_{i}}+Tn_{d_{i}}). Thus, the computational cost of the above formulation increases exponentially with the number of obstacles, time horizon, and dimension of the disturbances. In order to overcome this issue, we propose a second approach for deriving equivalent/approximate tractable versions of the constraint (12).

Reducing computational complexity

Let us first write the constraint (12) in terms of 𝜻i{\bm{\zeta}}^{i} as follows -

𝐀i𝝁xk,u¯i+𝐀i𝐌~ki𝜻i𝒃i2ci,𝜻iUi.\displaystyle\|{\bf A}_{i}{\bm{\mu}}_{x_{k},\bar{u}}^{i}+{\bf A}_{i}\tilde{{\bf M}}_{k}^{i}{\bm{\zeta}}^{i}-\bm{b}_{i}\|_{2}\geq c_{i},\quad\forall{\bm{\zeta}}^{i}\in\pazocal{U}_{i}. (29)

In this approach, we consider the following tighter approximation of the above constraint

𝐀i𝝁xk,u¯i𝒃i2𝐀i𝐌~ki𝜻i2ci,𝜻iUi,\displaystyle\|{\bf A}_{i}{\bm{\mu}}_{x_{k},\bar{u}}^{i}-\bm{b}_{i}\|_{2}-\|{\bf A}_{i}\tilde{{\bf M}}_{k}^{i}{\bm{\zeta}}^{i}\|_{2}\geq c_{i},\quad\forall{\bm{\zeta}}^{i}\in\pazocal{U}_{i}, (30)

which can be rewritten as

𝐀i𝝁xk,u¯i𝒃i2ci+max𝜻iUi𝐀i𝐌~ki𝜻i2.\displaystyle\|{\bf A}_{i}{\bm{\mu}}_{x_{k},\bar{u}}^{i}-\bm{b}_{i}\|_{2}\geq c_{i}+\max_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i}}\|{\bf A}_{i}\tilde{{\bf M}}_{k}^{i}{\bm{\zeta}}^{i}\|_{2}. (31)

Let us now introduce a slack variable c~i\tilde{c}_{i} such that we split the above constraint into the following system of two constraints.

𝐀i𝝁xk,u¯i𝒃i2c~i,\displaystyle\|{\bf A}_{i}{\bm{\mu}}_{x_{k},\bar{u}}^{i}-\bm{b}_{i}\|_{2}\geq\tilde{c}_{i}, (32)
max𝜻iUi𝐀i𝐌~ki𝜻i2c~ici\displaystyle\max_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i}}\|{\bf A}_{i}\tilde{{\bf M}}_{k}^{i}{\bm{\zeta}}^{i}\|_{2}\leq\tilde{c}_{i}-c_{i} (33)

Note however that the constraint (32) is non-convex. We address this by linearizing the constraint around the nominal trajectory 𝒖¯i,l\bar{{\bm{u}}}^{i,l}. In addition, the constraint (33) is still a semi-infinite constraint; thus, we need to derive its equivalent tractable constraints. Since the LHS in the constraint (33) involves a non-convex optimization problem, deriving the equivalent tractable constraints is NP-hard. Hence, we derive tighter approximations of the constraint (33) in the following proposition.

Proposition 2.

The tighter approximate constraints to the semi-infinite constraint (33) are given by

𝝁d,ki2c~ici\displaystyle\|{\bm{\mu}}_{d,k}^{i}\|_{2}\leq\tilde{c}_{i}-c_{i} (34)
(𝝁d,ki)m¯(τi)1/2𝚪iT𝐌iThk,m¯i𝐒i1\displaystyle({\bm{\mu}}_{d,k}^{i})_{\bar{m}}\geq(\tau^{i})^{1/2}||\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}h_{k,\bar{m}}^{i}||_{{\bf S}_{i}^{-1}} (35)

where hk,m¯iTh_{k,\bar{m}}^{i}{}^{\mathrm{T}} is m¯th\bar{m}^{th} row of 𝐀i𝐏ki{\bf A}_{i}{\bf P}_{k}^{i}.

Proof.

Provided in Section III of the SM. ∎

In this formulation, only the constraints (35) include 𝐒i{\bf S}_{i}. This formulation is useful in the case where there are multiple non-convex constraints of the form (12) at xkix_{k}^{i}, which have the same 𝐀i{\bf A}_{i} matrix. In such a case, we only need to enforce the constraint (35) once per each unique 𝐀i{\bf A}_{i} matrix in the constraints. For instance, let us again consider the collision avoidance example. Since the collision constraints (both obstacle and inter-agent) are applied to the agent’s position, the matrix 𝐀i{\bf A}_{i} remains the same in all such constraints, irrespective of the number of obstacles. Therefore, it is only the amount of constraints of the form (34) that would increase with the number of obstacles.

Subsequently, we proceed with considering the inter-agent robust non-convex constraints (13), which can be rewritten as

𝐀~ij𝔼[𝒙~kij]2cij,𝜻iUi,𝜻jUj,jNi,\displaystyle\|\tilde{{\bf A}}_{ij}\mathbb{E}[\tilde{{\bm{x}}}^{ij}_{k}]\|_{2}\geq c_{ij},\quad\forall{\bm{\zeta}}^{i}\in\pazocal{U}_{i},~{}{\bm{\zeta}}^{j}\in\pazocal{U}_{j},~{}j\in\pazocal{N}_{i}, (36)

where 𝐀~ij=[𝐀i,𝐀j]\tilde{{\bf A}}_{ij}=\begin{bmatrix}{\bf A}_{i},&-{\bf A}_{j}\end{bmatrix}, 𝒙~kij=[𝒙ki;𝒙kj]\tilde{{\bm{x}}}^{ij}_{k}=\begin{bmatrix}{\bm{x}}^{i}_{k};&{\bm{x}}^{j}_{k}\end{bmatrix}. The uncertainty involved for the variable 𝒙~ij\tilde{{\bm{x}}}^{ij} would be 𝜻~ij=[𝜻i;𝜻j]\tilde{{\bm{\zeta}}}^{ij}=\begin{bmatrix}{\bm{\zeta}}^{i};&{\bm{\zeta}}^{j}\end{bmatrix}. Further, it should be noted that if 𝜻i,𝜻j{\bm{\zeta}}^{i},{\bm{\zeta}}^{j} belong to the defined ellipsoidal sets Ui\pazocal{U}_{i}, Uj\pazocal{U}_{j} respectively, then the uncertainty vector 𝜻~ij\tilde{{\bm{\zeta}}}^{ij} can also be modeled by an ellipsoidal set of the defined form (II-D). Thus, the constraint (36) has a similar form as (12), and applying the initial approach would result in semi-definite constraints that are of the size n¯i+n¯j+1\bar{n}_{i}+\bar{n}_{j}+1. This might be particularly disadvantageous in a large-scale multi-agent setup, which might involve a significant number of neighbor agents. Thus, we derive equivalent/approximate tractable constraints using the second approach.

For that, let us first rewrite a single constraint corresponding to a neighbor agent jj of the agent ii (jNij\in\pazocal{N}_{i}) from the set of constraints (13) as follows.

𝐀i𝝁xk,u¯i𝐀j𝝁xk,u¯j+𝐀i𝐌~ki𝜻i𝐀j𝐌~kj𝜻j2cij,\displaystyle\|{\bf A}_{i}{\bm{\mu}}_{x_{k},\bar{u}}^{i}-{\bf A}_{j}{\bm{\mu}}_{x_{k},\bar{u}}^{j}+{\bf A}_{i}\tilde{{\bf M}}_{k}^{i}{\bm{\zeta}}^{i}-{\bf A}_{j}\tilde{{\bf M}}_{k}^{j}{\bm{\zeta}}^{j}\|_{2}\geq c_{ij},
𝜻iUi,𝜻jUj.\displaystyle~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}\forall{\bm{\zeta}}^{i}\in\pazocal{U}_{i},~{}{\bm{\zeta}}^{j}\in\pazocal{U}_{j}. (37)

We then consider the following tighter approximation to the above constraint

𝐀i𝝁xk,u¯i𝐀j𝝁xk,u¯j2𝐀i𝐌~ki𝜻i𝐀j𝐌~kj𝜻j2cij\displaystyle\|{\bf A}_{i}{\bm{\mu}}_{x_{k},\bar{u}}^{i}-{\bf A}_{j}{\bm{\mu}}_{x_{k},\bar{u}}^{j}\|_{2}-\|{\bf A}_{i}\tilde{{\bf M}}_{k}^{i}{\bm{\zeta}}^{i}-{\bf A}_{j}\tilde{{\bf M}}_{k}^{j}{\bm{\zeta}}^{j}\|_{2}\geq c_{ij}
𝜻iUi,𝜻jUj.\displaystyle~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}\forall{\bm{\zeta}}^{i}\in\pazocal{U}_{i},~{}{\bm{\zeta}}^{j}\in\pazocal{U}_{j}. (38)

and similarly with the previous case, we introduce a slack variable c~ij\tilde{c}_{ij}, through which we can rewrite the above constraint as follows

𝐀i𝝁xk,u¯i𝐀j𝝁xk,u¯j2c~ij,\displaystyle\|{\bf A}_{i}{\bm{\mu}}_{x_{k},\bar{u}}^{i}-{\bf A}_{j}{\bm{\mu}}_{x_{k},\bar{u}}^{j}\|_{2}\geq\tilde{c}_{ij}, (39)
max𝜻iUi,𝜻jUj\displaystyle\max_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i},~{}{\bm{\zeta}}^{j}\in\pazocal{U}_{j}} 𝐀i𝐌~ki𝜻i𝐀j𝐌~kj𝜻j2c~ijcij.\displaystyle\|{\bf A}_{i}\tilde{{\bf M}}_{k}^{i}{\bm{\zeta}}^{i}-{\bf A}_{j}\tilde{{\bf M}}_{k}^{j}{\bm{\zeta}}^{j}\|_{2}\leq\tilde{c}_{ij}-c_{ij}. (40)

The non-convex constraint (39) is addressed by linearizing it around the nominal value 𝒖¯i,l\bar{{\bm{u}}}^{i,l}. Similar to the previous case, the constraint (40) is semi-infinite, and deriving its equivalent tractable constraints is NP-hard. Thus, we derive the tighter approximations of the constraint (40).

Proposition 3.

The tighter approximate constraints to the semi-infinite constraint (40) are given by

𝝁d,ki+𝝁d,kj2c~ijcij,\displaystyle\|{\bm{\mu}}_{d,k}^{i}+{\bm{\mu}}_{d,k}^{j}\|_{2}\leq\tilde{c}_{ij}-c_{ij}, (41)
(𝝁d,ki)m¯(τi)1/2𝚪iT𝐌iThk,m¯i𝐒i1,\displaystyle({\bm{\mu}}_{d,k}^{i})_{\bar{m}}\geq(\tau^{i})^{1/2}||\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}h_{k,\bar{m}}^{i}||_{{\bf S}_{i}^{-1}}, (42)
(𝝁d,kj)m¯(τj)1/2𝚪jT𝐌jThk,m¯j𝐒j1,\displaystyle({\bm{\mu}}_{d,k}^{j})_{\bar{m}}\geq(\tau^{j})^{1/2}||\bm{\Gamma}_{j}^{\mathrm{T}}{\bf M}_{j}^{\mathrm{T}}h_{k,\bar{m}}^{j}||_{{\bf S}_{j}^{-1}}, (43)

where hk,m¯i,Thk,m¯jTh_{k,\bar{m}}^{i}{}^{\mathrm{T}},h_{k,\bar{m}}^{j}{}^{\mathrm{T}} are the m¯th\bar{m}^{th} rows of the matrices 𝐀i𝐏ki,𝐀j𝐏kj{\bf A}_{i}{\bf P}_{k}^{i},{\bf A}_{j}{\bf P}_{k}^{j} respectively.

It should be noted that the constraints (42), (43) will be the same as (35), if the 𝐀i{\bf A}_{i} matrix is the same (e.g., collision avoidance scenarios), therefore alleviating the need to introduce them as additional constraints in our framework.

III-A3 Robust linear chance constraints

The equivalent tractable constraints to the constraint (14) are given as [19]

𝒂iT𝝁x,u¯i+(τi)1/2𝚪iT𝐌iT𝒂i𝐒i1biα\displaystyle{\bm{a}}_{i}^{\mathrm{T}}{\bm{\mu}}_{x,\bar{u}}^{i}+(\tau^{i})^{1/2}||\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}{\bm{a}}_{i}||_{{\bf S}_{i}^{-1}}\leq b_{i}-\alpha (44)
η𝒂T(𝐆ui𝐊i+𝐈)𝐆𝒘i𝝋i2α\displaystyle\eta||{\bm{a}}^{\mathrm{T}}({\bf G}_{u}^{i}{\bf K}^{i}+{\bf I}){\bf G}_{{\bm{w}}}^{i}\bm{\varphi}^{i}||_{2}\leq\alpha (45)

where η:yN(0,1)(yq)=p\eta:\mathbb{P}_{y\in\pazocal{N}(0,1)}(y\geq q)=p, and 𝝋i𝝋i=T𝚺𝒘i\bm{\varphi}^{i}\bm{\varphi}^{i}{}^{\mathrm{T}}={\bm{\Sigma}}_{{\bm{w}}^{i}}.

III-A4 Robust covariance constraints

The covariance of the state of an agent ii is given as follows (derivation can be referenced from Section II of SM)

Cov[𝒙i]=(𝐆ui𝐊i+𝐈)𝐆𝒘i𝚺𝒘i𝐆𝒘i(𝐆ui𝐊i+𝐈)TT.\mathrm{Cov}[{\bm{x}}^{i}]=({\bf G}_{u}^{i}{\bf K}^{i}+{\bf I}){\bf G}_{{\bm{w}}}^{i}{\bm{\Sigma}}_{{\bm{w}}^{i}}{\bf G}_{{\bm{w}}}^{i}{}^{\mathrm{T}}({\bf G}_{u}^{i}{\bf K}^{i}+{\bf I})^{\mathrm{T}}. (46)

It should be observed that the covariance does not depend on the deterministic uncertainty but only on the covariance of the stochastic noise. Now, let us rewrite the constraint (15) in terms of 𝐊i{\bf K}^{i} as follows

(𝐆ui𝐊i+𝐈)𝐆𝒘i𝚺𝒘i𝐆𝒘i(𝐆ui𝐊i+𝐈)TT𝚺i.\displaystyle({\bf G}_{u}^{i}{\bf K}^{i}+{\bf I}){\bf G}_{{\bm{w}}}^{i}{\bm{\Sigma}}_{{\bm{w}}^{i}}{\bf G}_{{\bm{w}}}^{i}{}^{\mathrm{T}}({\bf G}_{u}^{i}{\bf K}^{i}+{\bf I}){}^{\mathrm{T}}\preceq{\bm{\Sigma}}^{i}. (47)

The above constraint is non-convex and can be reformulated into the following equivalent convex constraint by using the Schur complement,

[𝚺i(𝐆ui𝐊i+𝐈)𝐆𝒘i𝝋i((𝐆ui𝐊i+𝐈)𝐆𝒘i𝝋i)T𝐈(nx+Nnw)]0\begin{bmatrix}{\bm{\Sigma}}^{i}&({\bf G}_{u}^{i}{\bf K}^{i}+{\bf I}){\bf G}_{{\bm{w}}}^{i}\bm{\varphi}^{i}\\ (({\bf G}_{u}^{i}{\bf K}^{i}+{\bf I}){\bf G}_{{\bm{w}}}^{i}\bm{\varphi}^{i})^{\mathrm{T}}&{\bf I}_{(n_{x}+Nn_{w})}\end{bmatrix}\succeq 0 (48)

where 𝝋i𝝋i=T𝚺𝒘i\bm{\varphi}^{i}\bm{\varphi}^{i}{}^{\mathrm{T}}={\bm{\Sigma}}_{{\bm{w}}^{i}}.

III-B Problem Transformation

Now that we have derived the computationally tractable constraints that could be applied to a multi-robot setting, let us define the multi-agent optimization problem using the reformulated constraints.

Problem 2 (Multi-Agent Robust Optimization Problem II).

For all agents iVi\in\pazocal{V}, find the robust optimal policies 𝐮¯i,𝐊i\bar{{\bm{u}}}^{i},{\bf K}^{i}, such that

{𝒖¯i,𝐊i}iV=argminiVJi(𝒖¯i,𝐊i)\displaystyle\{\bar{{\bm{u}}}^{i},{\bf K}^{i}\}_{i\in\pazocal{V}}=\operatornamewithlimits{argmin}\sum_{i\in\pazocal{V}}J_{i}(\bar{{\bm{u}}}^{i},{\bf K}^{i})
s.t.\displaystyle\mathrm{s.t.}\; xk+1i=Akixki+Bkiuki+Ckidki+Dkiwki,k0,T1,\displaystyle~{}x_{k+1}^{i}=A_{k}^{i}x_{k}^{i}+B_{k}^{i}u_{k}^{i}+C_{k}^{i}d_{k}^{i}+D_{k}^{i}w_{k}^{i},~{}k\in\llbracket 0,T-1\rrbracket,
x¯0i : given,\displaystyle~{}\bar{x}_{0}^{i}\text{ : given}, (49)
(22),(24),(32),(34),(35),(39),(41),(42).\displaystyle\eqref{mean_constraint_form1},\eqref{New mean formulation},\eqref{nonconvex obstacle final ubar},\eqref{semi-definite obstacle new},\eqref{mu_d expression},\eqref{inter-agent collision avoidance ubar},\eqref{semi-definite interagent new},\eqref{interagent mu_d expression}.

We are interested in solving the above problem in a multi-robot setting. Hence, we consider the robust non-convex constraints to represent obstacle and inter-agent collision avoidance constraints and are applied at each time step kk. As mentioned in the previous subsection, all the collision avoidance constraints involve the same 𝐀i{\bf A}_{i} matrix in the constraint. Thus, in the subsequent sections, we consider 𝐀i{\bf A}_{i} in the constraints (32), (34), (35), (39), (41), (42) to be 𝐇pos{\bf H}_{pos}. Further, in this case, the constraints (35) and (42) are the same.

III-C Distributed Algorithm

Now that we have formulated the robust multi-agent trajectory optimization problem 2, we provide a distributed architecture for solving it.

III-C1 Problem Setup

In this subsection, we provide an overview of steps taken to reformulate the problem 2 into a structure that would allow us to solve it in a distributed manner. We start by considering the inter-agent constraints (39) and (41). To enforce these inter-agent constraints, each agent iVi\in\pazocal{V} needs {𝐇pos𝝁xk,u¯j,𝝁d,kj}k=0T\{{\bf H}_{pos}{\bm{\mu}}_{x_{k},\bar{u}}^{j},{\bm{\mu}}_{d,k}^{j}\}_{k=0}^{T} of each of its neighbor agents jNij\in\pazocal{N}_{i}. For convenience, let us assume 𝝁^xk,u¯i=𝐇pos𝝁xk,u¯i\hat{{\bm{\mu}}}_{x_{k},\bar{u}}^{i}={\bf H}_{pos}{\bm{\mu}}_{x_{k},\bar{u}}^{i} and define the sequences 𝝁^iu¯=[𝝁^x0,u¯i;𝝁^x1,u¯i;;𝝁^xT,u¯i]\hat{{\bm{\mu}}}_{i}^{\bar{u}}=\begin{bmatrix}\hat{{\bm{\mu}}}_{x_{0},\bar{u}}^{i};&\hat{{\bm{\mu}}}_{x_{1},\bar{u}}^{i};&\dots;\hat{{\bm{\mu}}}_{x_{T},\bar{u}}^{i}\end{bmatrix}, 𝝁id=[𝝁d,0i;𝝁d,1i;𝝁d,Ti]{\bm{\mu}}_{i}^{d}=\begin{bmatrix}{\bm{\mu}}_{d,0}^{i};&{\bm{\mu}}_{d,1}^{i};&\dots&{\bm{\mu}}_{d,T}^{i}\end{bmatrix}. We can rewrite problem 2 in a simplified way as

{𝒖¯i,𝐊i}iV=argmin\displaystyle\{\bar{{\bm{u}}}^{i},{\bf K}^{i}\}_{i\in\pazocal{V}}=\operatornamewithlimits{argmin} iVJ^i(𝒖¯i,𝐊i,𝝁^iu¯,𝝁id)\displaystyle\;\sum_{i\in\pazocal{V}}\hat{J}_{i}(\bar{{\bm{u}}}^{i},{\bf K}^{i},\hat{{\bm{\mu}}}_{i}^{\bar{u}},{\bm{\mu}}^{d}_{i})
s.t.iV,\displaystyle\mathrm{s.t.}\quad\forall\;i\in\pazocal{V},\; jNi,\displaystyle\forall\;j\in\pazocal{N}_{i},
Hi,j(𝝁^iu¯,𝝁^ju¯)0\displaystyle\pazocal{H}_{i,j}(\hat{{\bm{\mu}}}_{i}^{\bar{u}},\hat{{\bm{\mu}}}_{j}^{\bar{u}})\geq 0 (50a)
H~i,j(𝝁id,𝝁jd)0\displaystyle\tilde{\pazocal{H}}_{i,j}({\bm{\mu}}^{d}_{i},{\bm{\mu}}^{d}_{j})\geq 0 (50b)
where J^i(𝒖¯i,𝐊i,𝝁^iu¯,𝝁id)=Ji(𝒖¯i,𝐊i)+Ifi+Iu¯i+I𝐊i\displaystyle\hat{J}_{i}(\bar{{\bm{u}}}^{i},{\bf K}^{i},\hat{{\bm{\mu}}}_{i}^{\bar{u}},{\bm{\mu}}^{d}_{i})=J_{i}(\bar{{\bm{u}}}^{i},{\bf K}^{i})+\pazocal{I}_{f_{i}}+\pazocal{I}_{\bar{u}^{i}}+\pazocal{I}_{{\bf K}^{i}}
Ifi:Indicator function for (1)\displaystyle\pazocal{I}_{f_{i}}:\text{Indicator function for \eqref{dynamics}}
Iu¯i:Indicator function for (22), (24), (32)\displaystyle\pazocal{I}_{\bar{u}^{i}}:\text{Indicator function for \eqref{mean_constraint_form1}, \eqref{New mean formulation}, \eqref{nonconvex obstacle final ubar}}
I𝐊i:Indicator function for ((34), (35)\displaystyle\pazocal{I}_{{\bf K}^{i}}:\text{Indicator function for ((\ref{semi-definite obstacle new}), \eqref{mu_d expression}}
Hi,j(𝝁^iu¯,𝝁^ju¯)0:Constraint set representing (39)\displaystyle\pazocal{H}_{i,j}(\hat{{\bm{\mu}}}_{i}^{\bar{u}},\hat{{\bm{\mu}}}_{j}^{\bar{u}})\geq 0:\text{Constraint set representing \eqref{inter-agent collision avoidance ubar}}
H~i,j(𝝁id,𝝁jd)0:Constraint set representing (41).\displaystyle\tilde{\pazocal{H}}_{i,j}({\bm{\mu}}^{d}_{i},{\bm{\mu}}^{d}_{j})\geq 0:\text{Constraint set representing \eqref{semi-definite interagent new}}.

To solve the problem in a distributed manner, we first introduce the copy variables 𝝁¯ij,u¯,𝝁¯ij,d\bar{{\bm{\mu}}}_{i}^{j,\bar{u}},\bar{{\bm{\mu}}}_{i}^{j,d} at each agent ii corresponding to the variables 𝝁^ju¯,𝝁¯jd\hat{{\bm{\mu}}}^{\bar{u}}_{j},\bar{{\bm{\mu}}}^{d}_{j} of each of its neighbor agents jNij\in\pazocal{N}_{i}. Each agent then uses these copy variables to enforce the inter-agent constraints. However, these copy variables should capture the actual value of their corresponding variables. Thus, we need to enforce a consensus between the actual values of the variables 𝝁^iu¯,𝝁id\hat{{\bm{\mu}}}_{i}^{\bar{u}},{\bm{\mu}}_{i}^{d} for each agent ii and their corresponding copy variables 𝝁¯j^i,u¯\bar{{\bm{\mu}}}_{\hat{j}}^{i,\bar{u}}, 𝝁¯j^i,d\bar{{\bm{\mu}}}_{\hat{j}}^{i,d} respectively at each agent j^\hat{j} to which agent ii is a neighbor of (jPij\in\pazocal{P}_{i}). Hence, the above problem can be rewritten in terms of the copy variables as follows -

{𝒖¯i,𝐊i}iV=argminiVJ^i(𝒖¯i,𝐊i,𝝁^iu¯,𝝁id)\displaystyle\{\bar{{\bm{u}}}^{i},{\bf K}^{i}\}_{i\in\pazocal{V}}=\operatornamewithlimits{argmin}\;\sum_{i\in\pazocal{V}}\hat{J}_{i}(\bar{{\bm{u}}}^{i},{\bf K}^{i},\hat{{\bm{\mu}}}_{i}^{\bar{u}},{\bm{\mu}}^{d}_{i})
s.t.iV,\displaystyle\mathrm{s.t.}\quad\forall\;i\in\pazocal{V},
Hi,j(𝝁^iu¯,𝝁¯ij,u¯)0,jNi,\displaystyle\qquad\qquad\pazocal{H}_{i,j}(\hat{{\bm{\mu}}}_{i}^{\bar{u}},\bar{{\bm{\mu}}}_{i}^{j,\bar{u}})\geq 0,\;\forall\;j\in\pazocal{N}_{i}, (51a)
H~i,j(𝝁id,𝝁¯ij,d)0,jNi,\displaystyle\qquad\qquad\tilde{\pazocal{H}}_{i,j}({\bm{\mu}}_{i}^{d},\bar{{\bm{\mu}}}_{i}^{j,d})\geq 0,\;\forall\;j\in\pazocal{N}_{i}, (51b)
𝝁^iu¯=𝝁¯j^i,u¯,j^Pi\displaystyle\qquad\qquad\hat{{\bm{\mu}}}_{i}^{\bar{u}}=\bar{{\bm{\mu}}}_{\hat{j}}^{i,\bar{u}},\;\forall\;\hat{j}\in\pazocal{P}_{i} (51c)
𝝁id=𝝁¯j^i,d,j^Pi\displaystyle\qquad\qquad{\bm{\mu}}_{i}^{d}=\bar{{\bm{\mu}}}_{\hat{j}}^{i,d},\;\forall\;\hat{j}\in\pazocal{P}_{i} (51d)

Next, in order to use consensus ADMM (CADMM) [9], we introduce the global variables 𝝂iu¯,𝝂id{\bm{\nu}}_{i}^{\bar{u}},{\bm{\nu}}_{i}^{d} such that the constraints (51c), (51d) can be rewritten as follows.

𝝁^iu¯=𝝁¯j^i,u¯\displaystyle\hat{{\bm{\mu}}}_{i}^{\bar{u}}=\bar{{\bm{\mu}}}_{\hat{j}}^{i,\bar{u}} 𝝁^iu¯=𝝂iu¯,𝝂iu¯=𝝁¯j^i,u¯\displaystyle\iff\;\hat{{\bm{\mu}}}_{i}^{\bar{u}}={\bm{\nu}}_{i}^{\bar{u}},\;{\bm{\nu}}_{i}^{\bar{u}}=\bar{{\bm{\mu}}}_{\hat{j}}^{i,\bar{u}} (52)
𝝁id=𝝁¯j^i,d\displaystyle{\bm{\mu}}_{i}^{d}=\bar{{\bm{\mu}}}_{\hat{j}}^{i,d} 𝝁id=𝝂id,𝝁¯j^i,d=𝝂id\displaystyle\iff\;{\bm{\mu}}_{i}^{d}={\bm{\nu}}_{i}^{d},\;\bar{{\bm{\mu}}}_{\hat{j}}^{i,d}={\bm{\nu}}_{i}^{d} (53)

Then, we define the local variables of each agent ii as 𝝁~iu¯=[𝝁^iu¯;{𝝁¯ij,u¯}jNi]\tilde{{\bm{\mu}}}^{\bar{u}}_{i}=\begin{bmatrix}\hat{{\bm{\mu}}}_{i}^{\bar{u}};&\{\bar{{\bm{\mu}}}_{i}^{j,\bar{u}}\}_{j\in\pazocal{N}_{i}}\end{bmatrix}, 𝝁~id=[𝝁id;{𝝁¯ij,d}jNi]\tilde{{\bm{\mu}}}^{d}_{i}=\begin{bmatrix}{\bm{\mu}}_{i}^{d};&\{\bar{{\bm{\mu}}}_{i}^{j,d}\}_{j\in\pazocal{N}_{i}}\end{bmatrix}. For convenience, we also define the global variable sequences 𝝂~iu¯=[𝝂iu¯;{𝝂ju¯}jNi]\tilde{{\bm{\nu}}}_{i}^{\bar{u}}=\begin{bmatrix}{\bm{\nu}}_{i}^{\bar{u}};&\{{\bm{\nu}}_{j}^{\bar{u}}\}_{j\in\pazocal{N}_{i}}\end{bmatrix}, 𝝂~id=[𝝂id;{𝝂jd}jNi]\tilde{{\bm{\nu}}}_{i}^{d}=\begin{bmatrix}{\bm{\nu}}_{i}^{d};&\{{\bm{\nu}}_{j}^{d}\}_{j\in\pazocal{N}_{i}}\end{bmatrix}. Using the above-defined variables, problem (51) can be rewritten as

{𝒖¯i,𝐊i}iV\displaystyle\{\bar{{\bm{u}}}^{i},{\bf K}^{i}\}_{i\in\pazocal{V}} =argminiVJ^i(𝒖¯i,𝐊i,𝝁~iu¯,𝝁~id)\displaystyle=\operatornamewithlimits{argmin}\;\sum_{i\in\pazocal{V}}\hat{\pazocal{J}}_{i}(\bar{{\bm{u}}}^{i},{\bf K}^{i},\tilde{{\bm{\mu}}}^{\bar{u}}_{i},\tilde{{\bm{\mu}}}^{d}_{i}) (54)
s.t.iV,𝝁~iu¯=𝝂~iu¯,𝝁~id=𝝂~id\displaystyle\mathrm{s.t.}\;\forall\;i\in\pazocal{V},\quad\tilde{{\bm{\mu}}}^{\bar{u}}_{i}=\tilde{{\bm{\nu}}}_{i}^{\bar{u}},\quad\tilde{{\bm{\mu}}}^{d}_{i}=\tilde{{\bm{\nu}}}_{i}^{d}

where J^i(𝒖¯i,𝐊i,𝝁~iu¯,𝝁~id)=J^i(𝒖¯i,𝐊i,𝝁^iu¯,𝝁id)+IHi,j+IH~i,j\hat{\pazocal{J}}_{i}(\bar{{\bm{u}}}^{i},{\bf K}^{i},\tilde{{\bm{\mu}}}^{\bar{u}}_{i},\tilde{{\bm{\mu}}}^{d}_{i})=\hat{J}_{i}(\bar{{\bm{u}}}^{i},{\bf K}^{i},\hat{{\bm{\mu}}}_{i}^{\bar{u}},{\bm{\mu}}^{d}_{i})+\pazocal{I}_{\pazocal{H}_{i,j}}+\pazocal{I}_{\tilde{\pazocal{H}}_{i,j}}, and IHi,j\pazocal{I}_{\pazocal{H}_{i,j}}, IH~i,j\pazocal{I}_{\tilde{\pazocal{H}}_{i,j}} are indicator functions for the constraints (51a), (51b) respectively.

III-C2 Distributed Approach

Now that we have reformulated the problem to the form of (54), we can use CADMM to solve it in a distributed manner. The variables {𝝁~iu¯,𝝁~id}iV\{\tilde{{\bm{\mu}}}^{\bar{u}}_{i},\tilde{{\bm{\mu}}}^{d}_{i}\}_{i\in\pazocal{V}} constitute the first block, and {𝝂~iu¯,𝝂~id}i=1N\{\tilde{{\bm{\nu}}}^{\bar{u}}_{i},\tilde{{\bm{\nu}}}^{d}_{i}\}_{i=1}^{N} constitute the second block of the CADMM. In each iteration of CADMM, the Augmented Lagragian (AL) of the problem is minimized with respect to each block in a sequential manner followed by a dual update. The AL of the problem (54) is given by

Lρu¯,ρd=iV\displaystyle\pazocal{L}_{\rho_{\bar{u}},\rho_{d}}=\sum_{i\in\pazocal{V}} (J^i(𝒖¯i,𝐊i,𝝁~iu¯,𝝁~id)\displaystyle~{}\big{(}\hat{\pazocal{J}}_{i}(\bar{{\bm{u}}}^{i},{\bf K}^{i},\tilde{{\bm{\mu}}}^{\bar{u}}_{i},\tilde{{\bm{\mu}}}^{d}_{i}) (55)
+𝝀iu¯(𝝁~iu¯𝝂~iu¯)T+𝝀id(𝝁~id𝝂~id)T\displaystyle~{}~{}+{\bm{\lambda}}_{i}^{\bar{u}}{}^{\mathrm{T}}(\tilde{{\bm{\mu}}}^{\bar{u}}_{i}-\tilde{{\bm{\nu}}}^{\bar{u}}_{i})+{\bm{\lambda}}_{i}^{d}{}^{\mathrm{T}}(\tilde{{\bm{\mu}}}^{d}_{i}-\tilde{{\bm{\nu}}}^{d}_{i})
+ρu¯2||𝝁~iu¯𝝂~iu¯||22+ρd2||𝝁~id𝝂~id||22)\displaystyle~{}~{}+\frac{\rho_{\bar{u}}}{2}||\tilde{{\bm{\mu}}}^{\bar{u}}_{i}-\tilde{{\bm{\nu}}}^{\bar{u}}_{i}||_{2}^{2}+\frac{\rho_{d}}{2}||\tilde{{\bm{\mu}}}^{d}_{i}-\tilde{{\bm{\nu}}}^{d}_{i}||_{2}^{2}\big{)}

where 𝝀iu¯{\bm{\lambda}}_{i}^{\bar{u}}, 𝝀id{\bm{\lambda}}_{i}^{d} are dual variables, and ρu¯\rho_{\bar{u}}, ρd\rho_{d} are penalty parameters.

Refer to caption
(a) Robust case
Refer to caption
(b) Non-robust case
Figure 1: Two agents scenario with obstacle: Performance comparison between robust and non-robust case with terminal mean and obstacle constraints.

Now, the update steps in an lthl^{th} iteration of CADMM applied to the problem (54) are as follows.

Local variables update (Block - I)

The update step for the local variables is given by

{𝝁~iu¯,l+1,𝝁~id,l+1}iV\displaystyle\{\tilde{{\bm{\mu}}}^{\bar{u},l+1}_{i},\;\tilde{{\bm{\mu}}}^{d,l+1}_{i}\}_{i\in\pazocal{V}}
=argminLρu¯,ρd({𝝁~iu¯,𝝁~id,𝝂~iu¯,l,𝝂~id,l,𝝀iu¯,l,𝝀id,l}iV)\displaystyle~{}~{}~{}=\operatornamewithlimits{argmin}\;\pazocal{L}_{\rho_{\bar{u}},\rho_{d}}(\{\tilde{{\bm{\mu}}}^{\bar{u}}_{i},\tilde{{\bm{\mu}}}^{d}_{i},\tilde{{\bm{\nu}}}^{\bar{u},l}_{i},\tilde{{\bm{\nu}}}^{d,l}_{i},{\bm{\lambda}}_{i}^{\bar{u},l},{\bm{\lambda}}_{i}^{d,l}\}_{i\in\pazocal{V}})

Note that there is no coupling between the local variables of different agents. Thus, the above update step can be performed in a decentralized manner where each agent ii solves its sub-problem to update the local variables 𝝁~iu¯,l+1,𝝁~id,l+1\tilde{{\bm{\mu}}}^{\bar{u},l+1}_{i},\;\tilde{{\bm{\mu}}}^{d,l+1}_{i}.

Global variables update (Block - II)

Next, the global variables are updated through

{𝝂iu¯,l+1,𝝂id,l+1}iV\displaystyle\{{\bm{\nu}}^{\bar{u},l+1}_{i},\;{\bm{\nu}}^{d,l+1}_{i}\}_{i\in\pazocal{V}}
=argminLρu¯,ρd({𝝁~iu¯,l+1,𝝁~id,l+1,𝝂~iu¯,𝝂~id,𝝀iu¯,l,𝝀id,l}iV)\displaystyle~{}~{}~{}=\operatornamewithlimits{argmin}\;\pazocal{L}_{\rho_{\bar{u}},\rho_{d}}(\{\tilde{{\bm{\mu}}}^{\bar{u},l+1}_{i},\tilde{{\bm{\mu}}}^{d,l+1}_{i},\tilde{{\bm{\nu}}}^{\bar{u}}_{i},\tilde{{\bm{\nu}}}^{d}_{i},{\bm{\lambda}}_{i}^{\bar{u},l},{\bm{\lambda}}_{i}^{d,l}\}_{i\in\pazocal{V}})

Let us define 𝝀iu¯{\bm{\lambda}}_{i}^{\bar{u}}, 𝝀id{\bm{\lambda}}_{i}^{d} as sequence of dual vectors given as 𝝀iu¯=[𝝀¯iu¯;{𝝀¯ij,u¯}jNi]{\bm{\lambda}}_{i}^{\bar{u}}=\begin{bmatrix}\bar{{\bm{\lambda}}}_{i}^{\bar{u}};&\{\bar{{\bm{\lambda}}}_{i}^{j,\bar{u}}\}_{j\in\pazocal{N}_{i}}\end{bmatrix}, 𝝀iu¯=[𝝀¯id;{𝝀¯ij,d}jNi]{\bm{\lambda}}_{i}^{\bar{u}}=\begin{bmatrix}\bar{{\bm{\lambda}}}_{i}^{d};&\{\bar{{\bm{\lambda}}}_{i}^{j,d}\}_{j\in\pazocal{N}_{i}}\end{bmatrix}, such that the above update can be rewritten in a closed form for each agent iVi\in\pazocal{V} as follows (details are provided in Section III of the SM)

𝝂iu¯,l+1=1mi(𝝀¯iu¯,lρu¯+𝝁^iu¯,l+1+j^Pi(𝝀¯j^i,u¯,lρu¯+𝝁¯j^i,u¯,l+1))\displaystyle{\bm{\nu}}^{\bar{u},l+1}_{i}=\frac{1}{m_{i}}\bigg{(}\frac{\bar{{\bm{\lambda}}}_{i}^{\bar{u},l}}{\rho_{\bar{u}}}+\hat{{\bm{\mu}}}_{i}^{\bar{u},l+1}+\sum_{\hat{j}\in\pazocal{P}_{i}}\big{(}\frac{\bar{{\bm{\lambda}}}_{\hat{j}}^{i,\bar{u},l}}{\rho_{\bar{u}}}+\bar{{\bm{\mu}}}_{\hat{j}}^{i,\bar{u},l+1}\big{)}\bigg{)} (56)
𝝂id,l+1=1mi(𝝀¯id,lρd+𝝁id,l+1+j^Pi(𝝀¯j^i,d,lρd+𝝁¯j^i,d,l+1))\displaystyle{\bm{\nu}}^{d,l+1}_{i}=\frac{1}{m_{i}}\bigg{(}\frac{\bar{{\bm{\lambda}}}_{i}^{d,l}}{\rho_{d}}+{\bm{\mu}}^{d,l+1}_{i}+\sum_{\hat{j}\in\pazocal{P}_{i}}\big{(}\frac{\bar{{\bm{\lambda}}}_{\hat{j}}^{i,d,l}}{\rho_{d}}+\bar{{\bm{\mu}}}^{i,d,l+1}_{\hat{j}}\big{)}\bigg{)}

where mi=(n(Pi)+1)m_{i}=(n({\pazocal{P}_{i}})+1). After the local variables update, each agent ii would receive the updated variable 𝝁¯j^i,u¯,l+1\bar{{\bm{\mu}}}_{\hat{j}}^{i,\bar{u},l+1} from all the agents j^\hat{j} to which it is a neighbor of (j^Pi\hat{j}\in\pazocal{P}_{i}). Using these received local variables, each agent ii updates the global variables 𝝂iu¯,l+1,𝝂id,l+1{\bm{\nu}}^{\bar{u},l+1}_{i},{\bm{\nu}}^{d,l+1}_{i} as per the update step (56).

Dual variables update

Finally, the dual variables are updated through the following rules

𝝀iu¯,l+1=𝝀iu¯,l+ρu¯(𝝁~iu¯,l+1𝝂~iu¯,l+1)\displaystyle{\bm{\lambda}}_{i}^{\bar{u},l+1}={\bm{\lambda}}_{i}^{\bar{u},l}+\rho_{\bar{u}}(\tilde{{\bm{\mu}}}^{\bar{u},l+1}_{i}-\tilde{{\bm{\nu}}}^{\bar{u},l+1}_{i}) (57)
𝝀id,l+1=𝝀id,l+ρd(𝝁~id,l+1𝝂~id,l+1)\displaystyle{\bm{\lambda}}_{i}^{d,l+1}={\bm{\lambda}}_{i}^{d,l}+\rho_{d}(\tilde{{\bm{\mu}}}^{d,l+1}_{i}-\tilde{{\bm{\nu}}}^{d,l+1}_{i})

After the global variables update, each agent ii would receive the updated global variables 𝝂ju¯,l+1,𝝂jd,l+1{\bm{\nu}}^{\bar{u},l+1}_{j},{\bm{\nu}}^{d,l+1}_{j} from all its neighbor agents jNij\in\pazocal{N}_{i}. Using these received variables, the agent ii updates the dual variables 𝝀iu¯,l+1,𝝀id,l+1{\bm{\lambda}}_{i}^{\bar{u},l+1},{\bm{\lambda}}_{i}^{d,l+1} as per the update step (57). It should be noted that in each update step, all the agents do the update in parallel. Further, the agents need not exchange other system information (such as dynamics parameters or uncertainty set parameters).

IV Simulation Results

In this section, we provide simulation results of the proposed method applied to a multi-agent system with underlying stochastic and deterministic uncertainties. In Section IV-A, the effectiveness of the proposed robust constraints is analyzed under different scenarios. In Section IV-B, the scalability of the proposed distributed framework is illustrated. The agents have 2D double integrator dynamics with the exact dynamic matrices and the uncertainty information provided in Section IV of the SM. In all experiments, we set the time horizon to T=20T=20. All experiments are also illustrated through the supplementary video111https://youtu.be/yrDwPwg4WFI.

IV-A Effectiveness of Proposed Robust Constraints

We start by validating the effectiveness of the proposed robust constraints by considering different cases of underlying uncertainties. First, we analyze the constraints that are only on the expectation of the state. Subsequently, we analyze the robust chance constraints and covariance constraints, which are dependent on both the deterministic and stochastic disturbances.

IV-A1 With deterministic disturbances

Refer to caption
(a) Robust case
Refer to caption
(b) Non-robust case
Refer to caption
(c) Robust case, k=0k=0
Refer to caption
(d) Robust case, k=10k=10
Refer to caption
(e) Robust case, k=15k=15
Refer to caption
(f) Robust case, k=20k=20
Figure 2: Twenty agents scenario with circle formation task and obstacle: (a) with robust constraints, (b) without robust constraints, (c)-(f) Snapshots of mean trajectories of agents in robust case.

Here, we compare the results of applying the proposed distributed framework for two cases: robust and non-robust. With non-robust, we refer to the case of finding the optimal control trajectory by only considering standard optimization constraints instead of explicitly accounting for the deterministic uncertainties through robust optimization constraints (robust case). Fig. 1 discloses a two-agent scenario with terminal mean and obstacle avoidance constraints. The plots demonstrate 100100 trajectories of the two agents for different realizations of the disturbances. As shown in Fig. 1a, in the robust case, each agent ii successfully avoids the obstacle, and its disturbance-free state mean reaches the target terminal mean 𝝁targeti{\bm{\mu}}_{target}^{i} with uncertainty in its terminal state mean lying inside the set target bounds (represented by black boxes). On the contrary, for the non-robust case in Fig. 1b, the agents might collide with the obstacle in the center, and the uncertainty in the terminal state mean of each agent might lie outside the target bounds. In Fig. 2, we present a scenario of 20 agents in a circle formation, where each agent must reach the opposite side while avoiding collisions with other agents and an obstacle in the center. Fig. 2a and 2b show 100100 realizations for the robust and non-robust case. Clearly, the robust approach is able to successfully steer the agents to their targets while avoiding collisions, in contrast with the non-robust case where the agents might collide with the obstacle. In Fig. 2c - 2f, we show snapshots of the robust trajectories of the agents for different time instants, demonstrating that the agents do not violate any of the constraints during the tasks. Further, it is worth noting that in the robust case, the agents not only avoid collision but also the uncertainty in the state mean of an agent is reduced when it approaches an obstacle or another agent.

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Figure 3: Performance analysis of robust chance constraints and Covariance Constraints - Scenario 1: (a) Mean trajectories of agents in deterministic case, (b) Trajectories of agents in deterministic case without the robust chance constraints, (c) Trajectories of agents in mixed case with robust chance constraints, (d) Trajectories of agents in mixed case with robust chance constraints and terminal covariance constraints.

IV-A2 With both deterministic and stochastic disturbances

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Figure 4: Performance analysis of robust chance constraints - Scenario 2: (a) Mean trajectories of agents in the deterministic case, (b) Trajectories of agents in the deterministic case without the robust chance constraints, (c) Trajectories of agents in mixed case with robust chance constraints.

In this subsection, we demonstrate the effectiveness of the proposed framework with robust chance constraints and covariance constraints. For that, we compare two cases, one where the constraints that are only on the expectation of the state are active (‘deterministic case’), and another where all the robust constraints, including the robust chance and covariance constraints, are active (‘mixed case’). Further, it should be noted that when both disturbances are present, the mean trajectory differs from the actual trajectory.

We present two scenarios of a two-agent case with different levels of deterministic uncertainty controlled using τi\tau_{i} but with the same stochastic uncertainty characterized by initial state covariance to be diag([0.2,0.2,0.5,0.5])2\mathrm{diag}([0.2,0.2,0.5,0.5])^{2}, and covariance of each stochastic noise component wikw_{i}^{k} to be diag([0.02,0.02,0.2,0.2])2\mathrm{diag}([0.02,0.02,0.2,0.2])^{2} for each agent iVi\in\pazocal{V}. Fig. 3 discloses the first scenario with deterministic uncertainty parameterized by τ=0.01\tau=0.01 for both agents. It can be observed from Fig. 3a that the mean trajectories obtained by the proposed method in the deterministic case satisfy the inter-agent and the terminal mean constraints. However, Fig. 3b shows that in the presence of stochastic noise, a collision between the agents is highly probable, and the agents may not be able to reach the targets within the set bounds. Fig. 3c presents the case where we enforce robust chance constraints on the agents’ positions to stay within the blue dashed lines. As shown, most of the realizations of the trajectories lie within the specified bounds of the chance constraints, which ensures the safe operation of the agents despite the stochastic uncertainty. However, the uncertainty in the terminal state of the agents is not controlled by the applied robust chance constraints. We address this using covariance constraints. Fig 3d presents the scenario with additional terminal covariance constraints with the target covariance 𝚺i{\bm{\Sigma}}^{i} set to be diag([0.05,0.05,0.5,0.5])2\mathrm{diag}([0.05,0.05,0.5,0.5])^{2} for each agent iVi\in\pazocal{V}. It can be observed that the uncertainty in the terminal state is reduced.

Fig. 4 presents the second scenario of the two-agent case that involves a higher deterministic uncertainty parameterized by τ=0.05\tau=0.05 for both agents. Fig. 4b shows that the uncertainty in the terminal state of the agents might lie outside the target uncertainty bounds (indicated by black boxes). To address this, we can use covariance steering as shown in the previous scenario, or robust chance constraints. We use robust chance constraints in this scenario. Fig. 4c shows results for the mixed case when robust chance constraints are applied on the terminal state to stay within the target uncertainty bounds represented by the blue dashed box.

Further, it is interesting to note the following two observations. From Fig. 3a and 3d, it should be observed that uncertainty in the state trajectory is reduced to be smaller than the uncertainty in the mean in the deterministic case by adding covariance constraints. From Fig. 4a and 4b, it should be observed that the actual trajectories of the agents in the deterministic case avoid the obstacle without enforcing the robust chance constraints. This is because the feedback control gain 𝐊{\bf K} is used to control the deviations in the state due to the deterministic and stochastic uncertainties. Thus, controlling the deviation in the state due to one type of uncertainty also controls the deviation due to another. This might be useful in cases where one of the uncertainties (deterministic or stochastic) is significantly higher than the other. In such a case, we need not explicitly apply constraints to control the effect of the least pronounced uncertainty.

IV-B Scalability of Proposed Distributed Framework

In our experiments, we noticed that the centralized method with robust terminal mean and initial approach obstacle avoidance constraints alone would fail for a time horizon longer than 15 and a number of agents greater than 3. Fig. 5 shows the result of the proposed distributed framework applied to a large-scale scenario with 100 agents, which verifies the scalability of the proposed approach. The constraints considered are robust terminal mean and inter-agent constraints. Further, Fig. 6 reports the computational times of the proposed framework for an increasing number of agents for scenarios with robust terminal mean and inter-agent constraints. All simulations were carried out in Matlab2022b [17] environment using CVX [15] as the modeling software and MOSEK as the solver [2] on a system with an Intel Corei9-13900HX.

V Conclusion

In this paper, we merge principles from the areas of robust optimization, distribution steering and distributed optimization towards obtaining a scalable safe multi-agent control method under both stochastic and deterministic uncertainties. In particular, we present a distributed robust optimization method that exploits computationally tractable approximations of robust constraints and the separable nature of ADMM in order to significantly reduce the computational demands that such approaches would typically have when applied to multi-agent systems. The results verify that the proposed framework is able to effectively address both types of uncertainty and scale for multi-robot systems with up to a hundred agents.

In future work, we wish to extend the applicability of the proposed method and explore additional robust optimization formulations that would be relevant for robotics problems. In particular, since the current method considers linear time-varying dynamics, it would be possible to extend for nonlinear systems through sequential linearization schemes [32, 34]. Additionally, we intend to further explore suitable robust optimization formulations for multi-agent robotics and investigate connections with shooting methods [18, 41]. Finally, incorporating the current framework into hierarchical distributed optimization schemes [36] would be a promising direction for further increasing the scalability of the method.

Refer to caption
Figure 5: Large-scale scenario with 100 agents: Mean trajectories of the agents with robust terminal mean and inter-agent collision avoidance constraints.
Refer to caption
Figure 6: Scalability analysis: Computational times of the proposed distributed framework for an increasing number of agents.

Acknowledgments

This work was supported by the ARO Award #\#W911NF2010151. Augustinos Saravanos acknowledges financial support by the A. Onassis Foundation Scholarship.

References

  • Alvergue et al. [2016] Luis D Alvergue, Abhishek Pandey, Guoxiang Gu, and Xiang Chen. Consensus control for heterogeneous multiagent systems. SIAM Journal on Control and Optimization, 54(3):1719–1738, 2016.
  • ApS [2019] MOSEK ApS. The MOSEK optimization toolbox for MATLAB manual. Version 9.0., 2019. URL http://docs.mosek.com/9.0/toolbox/index.html.
  • Bakolas [2018] Efstathios Bakolas. Finite-horizon covariance control for discrete-time stochastic linear systems subject to input constraints. Automatica, 91:61–68, 2018.
  • Balci et al. [2022] Isin M Balci, Efstathios Bakolas, Bogdan Vlahov, and Evangelos A Theodorou. Constrained covariance steering based tube-mppi. In 2022 American Control Conference (ACC), pages 4197–4202. IEEE, 2022.
  • Ben-Tal et al. [2006a] Aharon Ben-Tal, Stephen Boyd, and Arkadi Nemirovski. Extending scope of robust optimization: Comprehensive robust counterparts of uncertain problems. Math. Program., 107:63–89, 06 2006a. doi: 10.1007/s10107-005-0679-z.
  • Ben-Tal et al. [2006b] Aharon Ben-Tal, Stephen Boyd, and Arkadi Nemirovski. Extending scope of robust optimization: Comprehensive robust counterparts of uncertain problems. Mathematical Programming, 107(1-2):63–89, 2006b.
  • Benedikter et al. [2022] Boris Benedikter, Alessandro Zavoli, Zhenbo Wang, Simone Pizzurro, and Enrico Cavallini. Convex approach to covariance control with application to stochastic low-thrust trajectory optimization. Journal of Guidance, Control, and Dynamics, 45(11):2061–2075, 2022.
  • Bertsimas et al. [2013] Dimitris Bertsimas, Eugene Litvinov, Xu Andy Sun, Jinye Zhao, and Tongxin Zheng. Adaptive robust optimization for the security constrained unit commitment problem. IEEE Transactions on Power Systems, 28(1):52–63, 2013. doi: 10.1109/TPWRS.2012.2205021.
  • Boyd et al. [2011] Stephen Boyd, Neal Parikh, Eric Chu, Borja Peleato, and Jonathan Eckstein. Distributed optimization and statistical learning via the alternating direction method of multipliers. Foundations and Trends® in Machine Learning, 3(1):1–122, 2011. ISSN 1935-8237. doi: 10.1561/2200000016.
  • Cai et al. [2020] Yuanxin Cai, Zhiqiang Wei, Ruide Li, Derrick Wing Kwan Ng, and Jinhong Yuan. Joint trajectory and resource allocation design for energy-efficient secure uav communication systems. IEEE Transactions on Communications, 68(7):4536–4553, 2020. doi: 10.1109/TCOMM.2020.2982152.
  • Chen et al. [2015] Yongxin Chen, Tryphon T Georgiou, and Michele Pavon. Optimal steering of a linear stochastic system to a final probability distribution, part i. IEEE Trans. Automat. Contr., 61(5):1158–1169, 2015.
  • Dorato [1987] Peter Dorato. A historical review of robust control. IEEE Control Systems Magazine, 7(2):44–47, 1987.
  • Feng et al. [2020] Zhi Feng, Guoqiang Hu, Yajuan Sun, and Jeffrey Soon. An overview of collaborative robotic manipulation in multi-robot systems. Annual Reviews in Control, 49:113–127, 2020.
  • Gómez et al. [2016] Vicenç Gómez, Sep Thijssen, Andrew Symington, Stephen Hailes, and Hilbert Kappen. Real-time stochastic optimal control for multi-agent quadrotor systems. In Proceedings of the International Conference on Automated Planning and Scheduling, volume 26, pages 468–476, 2016.
  • Grant and Boyd [2014] Michael Grant and Stephen Boyd. CVX: Matlab software for disciplined convex programming, version 2.1. http://cvxr.com/cvx, March 2014.
  • Hu et al. [2020] Junyan Hu, Parijat Bhowmick, Farshad Arvin, Alexander Lanzon, and Barry Lennox. Cooperative control of heterogeneous connected vehicle platoons: An adaptive leader-following approach. IEEE Robotics and Automation Letters, 5(2):977–984, 2020. doi: 10.1109/LRA.2020.2966412.
  • Inc. [2022] The MathWorks Inc. Matlab version: 9.13.0 (r2022b), 2022. URL https://www.mathworks.com.
  • Jacobson and Mayne [1970] David H Jacobson and David Q Mayne. Differential dynamic programming. Number 24. Elsevier Publishing Company, 1970.
  • Kotsalis et al. [2021] Georgios Kotsalis, Guanghui Lan, and Arkadi S. Nemirovski. Convex optimization for finite-horizon robust covariance control of linear stochastic systems. SIAM Journal on Control and Optimization, 59(1):296–319, 2021. doi: 10.1137/20M135090X. URL https://doi.org/10.1137/20M135090X.
  • Lee et al. [2022] Jaemin Lee, Efstathios Bakolas, and Luis Sentis. Hierarchical task-space optimal covariance control with chance constraints. IEEE Control Systems Letters, 6:2359–2364, 2022.
  • Lin et al. [2021] Yiheng Lin, Guannan Qu, Longbo Huang, and Adam Wierman. Multi-agent reinforcement learning in stochastic networked systems. Advances in neural information processing systems, 34:7825–7837, 2021.
  • Liu et al. [2022] Fengjiao Liu, George Rapakoulias, and Panagiotis Tsiotras. Optimal covariance steering for discrete-time linear stochastic systems. arXiv preprint arXiv:2211.00618, 2022.
  • Lorenzen et al. [2019] Matthias Lorenzen, Mark Cannon, and Frank Allgöwer. Robust mpc with recursive model update. Automatica, 103:461–471, 2019.
  • Nourian et al. [2012] Mojtaba Nourian, Peter E. Caines, Roland P. Malhame, and Minyi Huang. Mean field LQG control in leader-follower stochastic multi-agent systems: Likelihood ratio based adaptation. IEEE Transactions on Automatic Control, 57(11):2801–2816, 2012. doi: 10.1109/TAC.2012.2195797.
  • Okamoto and Tsiotras [2019] Kazuhide Okamoto and Panagiotis Tsiotras. Optimal stochastic vehicle path planning using covariance steering. IEEE Robotics and Automation Letters, 4(3):2276–2281, 2019. doi: 10.1109/LRA.2019.2901546.
  • Pereira et al. [2022] Marcus A Pereira, Augustinos D Saravanos, Oswin So, and Evangelos A. Theodorou. Decentralized Safe Multi-agent Stochastic Optimal Control using Deep FBSDEs and ADMM. In Proceedings of Robotics: Science and Systems, New York City, NY, USA, June 2022. doi: 10.15607/RSS.2022.XVIII.055.
  • Petersen et al. [2000] Ian R. Petersen, Valery A. Ugrinovskii, and Andrey V. Savkin. Introduction, pages 1–18. Springer London, London, 2000. ISBN 978-1-4471-0447-6. doi: 10.1007/978-1-4471-0447-6˙1. URL https://doi.org/10.1007/978-1-4471-0447-6_1.
  • Pilipovsky and Tsiotras [2023] Joshua Pilipovsky and Panagiotis Tsiotras. Data-driven robust covariance control for uncertain linear systems. arXiv preprint arXiv:2312.05833, 2023.
  • Pishvaee et al. [2011] Mir Saman Pishvaee, Masoud Rabbani, and Seyed Ali Torabi. A robust optimization approach to closed-loop supply chain network design under uncertainty. Applied mathematical modelling, 35(2):637–649, 2011.
  • Quintero-Pena et al. [2021] Carlos Quintero-Pena, Anastasios Kyrillidis, and Lydia E Kavraki. Robust optimization-based motion planning for high-dof robots under sensing uncertainty. In 2021 IEEE International Conference on Robotics and Automation (ICRA), pages 9724–9730. IEEE, 2021.
  • Rapakoulias and Tsiotras [2023] George Rapakoulias and Panagiotis Tsiotras. Discrete-time optimal covariance steering via semidefinite programming. arXiv preprint arXiv:2302.14296, 2023.
  • Ridderhof et al. [2019] Jack Ridderhof, Kazuhide Okamoto, and Panagiotis Tsiotras. Nonlinear uncertainty control with iterative covariance steering. In 2019 IEEE 58th Conference on Decision and Control (CDC), pages 3484–3490, 2019. doi: 10.1109/CDC40024.2019.9029993.
  • Saravanos et al. [2021] Augustinos D Saravanos, Alexandros Tsolovikos, Efstathios Bakolas, and Evangelos Theodorou. Distributed Covariance Steering with Consensus ADMM for Stochastic Multi-Agent Systems. In Proceedings of Robotics: Science and Systems, Virtual, July 2021. doi: 10.15607/RSS.2021.XVII.075.
  • Saravanos et al. [2022] Augustinos D Saravanos, Isin M Balci, Efstathios Bakolas, and Evangelos A Theodorou. Distributed model predictive covariance steering. arXiv preprint arXiv:2212.00398, 2022.
  • Saravanos et al. [2023a] Augustinos D. Saravanos, Yuichiro Aoyama, Hongchang Zhu, and Evangelos A. Theodorou. Distributed differential dynamic programming architectures for large-scale multiagent control. IEEE Transactions on Robotics, 39(6):4387–4407, 2023a. doi: 10.1109/TRO.2023.3319894.
  • Saravanos et al. [2023b] Augustinos D Saravanos, Yihui Li, and Evangelos Theodorou. Distributed Hierarchical Distribution Control for Very-Large-Scale Clustered Multi-Agent Systems. In Proceedings of Robotics: Science and Systems, Daegu, Republic of Korea, July 2023b. doi: 10.15607/RSS.2023.XIX.110.
  • Schranz et al. [2020] Melanie Schranz, Martina Umlauft, Micha Sende, and Wilfried Elmenreich. Swarm robotic behaviors and current applications. Frontiers in Robotics and AI, page 36, 2020.
  • Shah et al. [2022] Kunal Shah, Annie Schmidt, Grant Ballard, and Mac Schwager. Large scale aerial multi-robot coverage path planning. Field Robotics, 2(1), 2022.
  • Shamma [2008] Jeff Shamma. Cooperative control of distributed multi-agent systems. John Wiley & Sons, 2008.
  • Taylor et al. [2021] Andrew J Taylor, Victor D Dorobantu, Sarah Dean, Benjamin Recht, Yisong Yue, and Aaron D Ames. Towards robust data-driven control synthesis for nonlinear systems with actuation uncertainty. In 2021 60th IEEE Conference on Decision and Control (CDC), pages 6469–6476. IEEE, 2021.
  • Von Stryk and Bulirsch [1992] Oskar Von Stryk and Roland Bulirsch. Direct and indirect methods for trajectory optimization. Annals of operations research, 37(1):357–373, 1992.
  • Wan et al. [2021] Neng Wan, Aditya Gahlawat, Naira Hovakimyan, Evangelos A. Theodorou, and Petros G. Voulgaris. Cooperative path integral control for stochastic multi-agent systems. In 2021 American Control Conference (ACC), pages 1262–1267, 2021. doi: 10.23919/ACC50511.2021.9482942.
  • Xiao et al. [2021] Wei Xiao, Christos G Cassandras, and Calin A Belta. Bridging the gap between optimal trajectory planning and safety-critical control with applications to autonomous vehicles. Automatica, 129:109592, 2021.
  • Yin et al. [2022] Ji Yin, Zhiyuan Zhang, Evangelos Theodorou, and Panagiotis Tsiotras. Trajectory distribution control for model predictive path integral control using covariance steering. In 2022 International Conference on Robotics and Automation (ICRA), pages 1478–1484. IEEE, 2022.
  • Zhou and Doyle [1998] Kemin Zhou and John Comstock Doyle. Essentials of robust control, volume 104. Prentice hall Upper Saddle River, NJ, 1998.
  • Zhu et al. [2021] Pingping Zhu, Chang Liu, and Silvia Ferrari. Adaptive online distributed optimal control of very-large-scale robotic systems. IEEE Transactions on Control of Network Systems, 8(2):678–689, 2021. doi: 10.1109/TCNS.2021.3097306.

Supplementary Material

In the following, we provide detailed derivations, proofs and experiment data that have been omitted in the main paper.

VI System Dynamics Compact Form

The system dynamics (1) can be written in a more compact form as follows

𝒙i=𝐆0ix¯0i+𝐆ui𝒖i+𝐆𝒘i𝒘i+𝐆ζi𝜻i,{\bm{x}}^{i}={\bf G}_{0}^{i}\bar{x}_{0}^{i}+{\bf G}_{u}^{i}{\bm{u}}^{i}+{\bf G}_{{\bm{w}}}^{i}\bm{{\bm{w}}}^{i}+{\bf G}_{\zeta}^{i}\bm{\zeta}^{i}, (58)

where

𝐆0i=[𝐈;𝚽i(1,0);𝚽i(2,0);;𝚽i(T,0)],\displaystyle{\bf G}_{0}^{i}=\begin{bmatrix}{\bf I};&{\bf\Phi}^{i}(1,0);&{\bf\Phi}^{i}(2,0);&\dots;&{\bf\Phi}^{i}(T,0)\end{bmatrix}, (59a)
𝐆ui=[𝟎𝟎𝟎B0i𝟎𝟎𝚽i(2,1)B0iB1i𝟎𝚽i(T,1)B0i𝚽i(T,2)B1iBT1i],\displaystyle{\bf G}_{u}^{i}=\begin{bmatrix}{\bf 0}&{\bf 0}&\dots&{\bf 0}\\ B_{0}^{i}&{\bf 0}&\dots&{\bf 0}\\ {\bf\Phi}^{i}(2,1)B_{0}^{i}&B_{1}^{i}&\dots&{\bf 0}\\ \vdots&\vdots&\vdots&\vdots\\ {\bf\Phi}^{i}(T,1)B_{0}^{i}&{\bf\Phi}^{i}(T,2)B_{1}^{i}&\dots&B_{T-1}^{i}\end{bmatrix}, (59b)
𝐆𝒘i=[𝐈𝟎𝟎𝚽i(1,0)D0i𝟎𝚽i(2,0)𝚽i(2,1)D0i𝟎𝚽i(T,0)𝚽i(T,1)D0iDT1i],\displaystyle{\bf G}_{{\bm{w}}}^{i}=\begin{bmatrix}{\bf I}&{\bf 0}&\dots&{\bf 0}\\ {\bf\Phi}^{i}(1,0)&D_{0}^{i}&\dots&{\bf 0}\\ {\bf\Phi}^{i}(2,0)&{\bf\Phi}^{i}(2,1)D_{0}^{i}&\dots&{\bf 0}\\ \vdots&\vdots&\vdots&\vdots\\ {\bf\Phi}^{i}(T,0)&{\bf\Phi}^{i}(T,1)D_{0}^{i}&\dots&D_{T-1}^{i}\end{bmatrix}, (59c)
𝐆ζi=[𝐈𝟎𝟎𝚽i(1,0)C0i𝟎𝚽i(2,0)𝚽i(2,1)C0i𝟎𝚽i(T,0)𝚽i(T,1)C0iCT1i],\displaystyle{\bf G}_{\zeta}^{i}=\begin{bmatrix}{\bf I}&{\bf 0}&\dots&{\bf 0}\\ {\bf\Phi}^{i}(1,0)&C_{0}^{i}&\dots&{\bf 0}\\ {\bf\Phi}^{i}(2,0)&{\bf\Phi}^{i}(2,1)C_{0}^{i}&\dots&{\bf 0}\\ \vdots&\vdots&\vdots&\vdots\\ {\bf\Phi}^{i}(T,0)&{\bf\Phi}^{i}(T,1)C_{0}^{i}&\dots&C_{T-1}^{i}\end{bmatrix}, (59d)

with

𝚽i(k1,k2)=Ak11iAk12iAk2ifork1>k2.\displaystyle{\bf\Phi}^{i}(k_{1},k_{2})=A_{k_{1}-1}^{i}A_{k_{1}-2}^{i}\dots A_{k_{2}}^{i}\;\text{for}\;k_{1}>k_{2}. (60)

VII Robust Linear Constraints

VII-A Expectation of State

Let us rewrite the state equation for an agent ii as follows

𝒙i=𝐆0ix¯0i+𝐆ui𝒖i+𝐆𝒘i𝒘i+𝐆ζi𝜻i\displaystyle{\bm{x}}^{i}={\bf G}_{0}^{i}\bar{x}_{0}^{i}+{\bf G}_{u}^{i}{\bm{u}}^{i}+{\bf G}_{{\bm{w}}}^{i}\bm{{\bm{w}}}^{i}+{\bf G}_{\zeta}^{i}\bm{\zeta}^{i}

Substituting the control 𝒖i{\bm{u}}^{i} as 𝒖¯i+𝐊i𝜹i\bar{{\bm{u}}}^{i}+{\bf K}^{i}{\bm{\delta}}^{i} (from (10)) in the above state equation, we get the following.

𝒙i=𝐆0ix¯0i+𝐆ui(𝒖¯i+𝐊i𝜹i)+𝐆𝒘i𝒘i+𝐆ζi𝜻i\displaystyle{\bm{x}}^{i}={\bf G}_{0}^{i}\bar{x}_{0}^{i}+{\bf G}_{u}^{i}\big{(}\bar{{\bm{u}}}^{i}+{\bf K}^{i}{\bm{\delta}}^{i}\big{)}+{\bf G}_{{\bm{w}}}^{i}\bm{{\bm{w}}}^{i}+{\bf G}_{\zeta}^{i}\bm{\zeta}^{i} (61)

Further, using (8), we can rewrite the above equation as follows

𝒙i\displaystyle{\bm{x}}^{i} =𝐆0ix¯0i+𝐆ui(𝒖¯i+𝐊i(𝐆𝒘i𝒘i+𝐆ζi𝜻i))\displaystyle={\bf G}_{0}^{i}\bar{x}_{0}^{i}+{\bf G}_{u}^{i}\big{(}\bar{{\bm{u}}}^{i}+{\bf K}^{i}({\bf G}_{{\bm{w}}}^{i}\bm{{\bm{w}}}^{i}+{\bf G}_{\zeta}^{i}\bm{\zeta}^{i})\big{)} (62)
+𝐆𝒘i𝒘i+𝐆ζi𝜻i\displaystyle~{}~{}+{\bf G}_{{\bm{w}}}^{i}\bm{{\bm{w}}}^{i}+{\bf G}_{\zeta}^{i}\bm{\zeta}^{i}

which can be written in compact form as follows -

𝒙i\displaystyle{\bm{x}}^{i} =𝐆0ix¯0i+𝐆ui𝒖¯i+(𝐆ui𝐊i+𝐈)𝐆𝒘i𝒘i\displaystyle={\bf G}_{0}^{i}\bar{x}_{0}^{i}+{\bf G}_{u}^{i}\bar{{\bm{u}}}^{i}+({\bf G}_{u}^{i}{\bf K}^{i}+{\bf I}){\bf G}_{{\bm{w}}}^{i}\bm{{\bm{w}}}^{i} (63)
+(𝐆ui𝐊i+𝐈)𝐆ζi𝜻i\displaystyle~{}~{}~{}+({\bf G}_{u}^{i}{\bf K}^{i}+{\bf I}){\bf G}_{\zeta}^{i}\bm{\zeta}^{i}

Now, the expectation of the state can be given as

E[𝒙i]\displaystyle E[{\bm{x}}^{i}] =𝐆0ix¯0i+𝐆ui𝒖¯i+(𝐆ui𝐊i+𝐈)𝐆𝒘iE[𝒘i]\displaystyle={\bf G}_{0}^{i}\bar{x}_{0}^{i}+{\bf G}_{u}^{i}\bar{{\bm{u}}}^{i}+({\bf G}_{u}^{i}{\bf K}^{i}+{\bf I}){\bf G}_{{\bm{w}}}^{i}E[\bm{{\bm{w}}}^{i}] (64)
+(𝐆ui𝐊i+𝐈)𝐆ζi𝜻i\displaystyle~{}~{}~{}+({\bf G}_{u}^{i}{\bf K}^{i}+{\bf I}){\bf G}_{\zeta}^{i}\bm{\zeta}^{i}

Finally, using the fact that E[𝒘i]=0E[\bm{{\bm{w}}}^{i}]=0, we get

E[𝒙i]=𝐆0ix¯0i+𝐆ui𝒖¯i+(𝐆ui𝐊i+𝐈)𝐆ζi𝜻iE[{\bm{x}}^{i}]={\bf G}_{0}^{i}\bar{x}_{0}^{i}+{\bf G}_{u}^{i}\bar{{\bm{u}}}^{i}+({\bf G}_{u}^{i}{\bf K}^{i}+{\bf I}){\bf G}_{\zeta}^{i}\bm{\zeta}^{i} (65)

VII-B Covariance of State

The covariance of the state of an agent ii is given as follows

Cov[𝒙i]=𝔼[(𝒙i𝔼[𝒙i])(𝒙i𝔼[𝒙i])T]\mathrm{Cov}[{\bm{x}}^{i}]=\mathbb{E}\Big{[}({\bm{x}}^{i}-\mathbb{E}[{\bm{x}}^{i}])({\bm{x}}^{i}-\mathbb{E}[{\bm{x}}^{i}])^{\mathrm{T}}\Big{]} (66)

Let us first write the expression for 𝒙i𝔼[𝒙i]{\bm{x}}^{i}-\mathbb{E}[{\bm{x}}^{i}] using (63) and (65) as follows

𝒙i𝔼[𝒙i]=(𝐆ui𝐊i+𝐈)𝐆𝒘i𝒘i{\bm{x}}^{i}-\mathbb{E}[{\bm{x}}^{i}]=({\bf G}_{u}^{i}{\bf K}^{i}+{\bf I}){\bf G}_{{\bm{w}}}^{i}\bm{{\bm{w}}}^{i} (67)

Now, the covariance of the state can be rewritten as -

Cov[𝒙i]=𝔼[(𝐆ui𝐊i+𝐈)𝐆𝒘i𝒘i((𝐆ui𝐊i+𝐈)𝐆𝒘i𝒘i)T]\mathrm{Cov}[{\bm{x}}^{i}]=\mathbb{E}\Big{[}({\bf G}_{u}^{i}{\bf K}^{i}+{\bf I}){\bf G}_{{\bm{w}}}^{i}\bm{{\bm{w}}}^{i}(({\bf G}_{u}^{i}{\bf K}^{i}+{\bf I}){\bf G}_{{\bm{w}}}^{i}\bm{{\bm{w}}}^{i})^{\mathrm{T}}\Big{]} (68)

which can be further rewritten as

Cov[𝒙i]=(𝐆ui𝐊i+𝐈)𝐆𝒘i𝔼[𝒘i𝒘i]T𝐆𝒘i(𝐆ui𝐊i+𝐈)TT\mathrm{Cov}[{\bm{x}}^{i}]=({\bf G}_{u}^{i}{\bf K}^{i}+{\bf I}){\bf G}_{{\bm{w}}}^{i}\mathbb{E}[\bm{{\bm{w}}}^{i}\bm{{\bm{w}}}^{i}{}^{\mathrm{T}}]{\bf G}_{{\bm{w}}}^{i}{}^{\mathrm{T}}({\bf G}_{u}^{i}{\bf K}^{i}+{\bf I})^{\mathrm{T}} (69)

Finally, using the fact that 𝔼[𝒘i𝒘i]T=𝚺𝒘i\mathbb{E}[\bm{{\bm{w}}}^{i}\bm{{\bm{w}}}^{i}{}^{\mathrm{T}}]={\bm{\Sigma}}_{{\bm{w}}^{i}}, we get

Cov[𝒙i]=(𝐆ui𝐊i+𝐈)𝐆𝒘i𝚺𝒘i𝐆𝒘i(𝐆ui𝐊i+𝐈)TT\mathrm{Cov}[{\bm{x}}^{i}]=({\bf G}_{u}^{i}{\bf K}^{i}+{\bf I}){\bf G}_{{\bm{w}}}^{i}{\bm{\Sigma}}_{{\bm{w}}^{i}}{\bf G}_{{\bm{w}}}^{i}{}^{\mathrm{T}}({\bf G}_{u}^{i}{\bf K}^{i}+{\bf I})^{\mathrm{T}} (70)

VII-C Proposition 1 Proof

The expression max𝜻iUi𝒂iT𝔼[𝒙i]\max_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i}}{\bm{a}}_{i}^{\mathrm{T}}\mathbb{E}[{\bm{x}}^{i}] can be rewritten as

max𝜻iUi𝒂iT𝔼[𝒙i]\displaystyle\max_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i}}{\bm{a}}_{i}^{\mathrm{T}}\mathbb{E}[{\bm{x}}^{i}] =𝒂iT𝝁x,u¯i+max𝜻iUi𝒂iT𝐌i𝜻i\displaystyle={\bm{a}}_{i}^{\mathrm{T}}{\bm{\mu}}_{x,\bar{u}}^{i}+\max_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i}}{\bm{a}}_{i}^{\mathrm{T}}{\bf M}_{i}{\bm{\zeta}}^{i} (71a)
=𝒂iT𝝁x,u¯imin𝐒i𝒛i,𝒛iτi𝒂iT𝐌i𝚪i𝒛i.\displaystyle={\bm{a}}_{i}^{\mathrm{T}}{\bm{\mu}}_{x,\bar{u}}^{i}-\min_{\langle{\bf S}_{i}{\bm{z}}_{i},{\bm{z}}_{i}\rangle\leq\tau^{i}}-{\bm{a}}_{i}^{\mathrm{T}}{\bf M}_{i}\bm{\Gamma}_{i}{\bm{z}}_{i}. (71b)

Let us now consider the Lagrangian for the minimization problem inside (71b), given by

L(𝐳i,λ)=𝐚iT𝐌i𝚪i𝐳i+λ(𝐒i𝐳i,𝐳iτi),\displaystyle\pazocal{L}({\bm{z}}_{i},\lambda)=-{\bm{a}}_{i}^{\mathrm{T}}{\bf M}_{i}\bm{\Gamma}_{i}{\bm{z}}_{i}+\lambda(\langle{\bf S}_{i}{\bm{z}}_{i},{\bm{z}}_{i}\rangle-\tau^{i}), (72)

where λ\lambda\in\mathbb{R} is the Lagrange multiplier corresponding to the inequality constraint 𝐒i𝒛i,𝒛iτi\langle{\bf S}_{i}{\bm{z}}_{i},{\bm{z}}_{i}\rangle\leq\tau^{i}. The above problem is convex, and the constraint set 𝐒i𝒛i,𝒛iτi\langle{\bf S}_{i}{\bm{z}}_{i},{\bm{z}}_{i}\rangle\leq\tau^{i} is non-empty - since 𝒛i=0{\bm{z}}_{i}=0 lies in the interior of the feasible set because τi>0\tau^{i}>0. Thus, the problem satisfies Slater’s condition, and we could find the minimizer using KKT conditions as follows

𝒛iL(𝐳i,λ)=0,\displaystyle\nabla_{{\bm{z}}_{i}}\pazocal{L}({\bm{z}}_{i}^{*},\lambda^{*})=0, (73a)
λ(𝐒i𝒛i,𝒛iτi)=0,\displaystyle\lambda^{*}(\langle{\bf S}_{i}{\bm{z}}_{i}^{*},{\bm{z}}_{i}^{*}\rangle-\tau^{i})=0, (73b)
λ0,\displaystyle\lambda^{*}\geq 0, (73c)
𝐒i𝒛i,𝒛iτi0.\displaystyle\langle{\bf S}_{i}{\bm{z}}_{i}^{*},{\bm{z}}_{i}^{*}\rangle-\tau^{i}\leq 0. (73d)

Let us first simplify the condition 𝒛iL(𝐳i,λ)=0\nabla_{{\bm{z}}_{i}}\pazocal{L}({\bm{z}}_{i}^{*},\lambda^{*})=0, as

𝚪iT𝐌iT𝒂i+2λ𝐒i𝒛i=0,-\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}{\bm{a}}_{i}+2\lambda^{*}{\bf S}_{i}{\bm{z}}_{i}^{*}=0, (74)

which leads to

2λ𝒛i=𝐒i1𝚪iT𝐌iT𝒂i.2\lambda^{*}{\bm{z}}_{i}^{*}={\bf S}_{i}^{-1}\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}{\bm{a}}_{i}. (75)

It can be observed that 𝚪iT𝐌iT𝒂i\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}{\bm{a}}_{i} cannot be equal to the zero vector since it is the coefficient of 𝒛i{\bm{z}}_{i} in the objective function. Thus, since we also have that 𝐒i𝕊++n¯i{\bf S}_{i}\in\mathbb{S}_{++}^{\bar{n}_{i}}, then the RHS of the above equation is non-zero, which implies that both 𝒛i{\bm{z}}_{i}^{*} and λ\lambda^{*} need to be non-zero. Therefore, we obtain

𝒛i=(2λ)1𝐒i1𝚪iT𝐌iT𝒂i.{\bm{z}}_{i}^{*}=(2\lambda^{*})^{-1}{\bf S}_{i}^{-1}\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}{\bm{a}}_{i}. (76)

Further, from the condition (73b) and using λ0\lambda^{*}\neq 0, we get

𝐒i𝒛i,𝒛iτi=0,\langle{\bf S}_{i}{\bm{z}}_{i}^{*},{\bm{z}}_{i}^{*}\rangle-\tau^{i}=0, (77)

which through (76) gives

(2λ)2𝚪iT𝐌iT𝒂i,𝐒i1𝚪iT𝐌iT𝒂i=τi,(2\lambda^{*})^{-2}\langle\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}{\bm{a}}_{i},\;{\bf S}_{i}^{-1}\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}{\bm{a}}_{i}\rangle=\tau^{i}, (78)

The above relationship leads to

(τi)1/2𝚪iT𝐌iT𝒂i𝐒i1=2λ,(\tau^{i})^{-1/2}\;\|\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}{\bm{a}}_{i}\|_{{\bf S}_{i}^{-1}}=2\lambda^{*}, (79)

since λ>0\lambda^{*}>0. The optimal value L(𝐳i,λ)\pazocal{L}({\bm{z}}_{i}^{*},\lambda^{*}) is then given by

L(𝐳i,λ)\displaystyle\pazocal{L}({\bm{z}}_{i}^{*},\lambda^{*}) =𝒂iT𝐌i𝚪i𝒛i\displaystyle=-{\bm{a}}_{i}^{\mathrm{T}}{\bf M}_{i}\bm{\Gamma}_{i}{\bm{z}}_{i}^{*} (80a)
=(2λ)1𝒂iT𝐌i𝚪i𝐒i1𝚪iT𝐌iT𝒂i\displaystyle=-(2\lambda^{*})^{-1}{\bm{a}}_{i}^{\mathrm{T}}{\bf M}_{i}\bm{\Gamma}_{i}{\bf S}_{i}^{-1}\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}{\bm{a}}_{i} (80b)
=(2λ)1𝚪iT𝐌iT𝒂i𝐒i12\displaystyle=-(2\lambda^{*})^{-1}\|\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}{\bm{a}}_{i}\|_{{\bf S}_{i}^{-1}}^{2} (80c)
=(τi)1/2𝚪iT𝐌iT𝒂i𝐒i1.\displaystyle=-(\tau^{i})^{1/2}\|\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}{\bm{a}}_{i}\|_{{\bf S}_{i}^{-1}}. (80d)

Therefore, we obtain

max𝜻iUi𝒂iT𝔼[𝒙i]\displaystyle\max_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i}}{\bm{a}}_{i}^{\mathrm{T}}\mathbb{E}[{\bm{x}}^{i}] =𝒂iT𝝁x,u¯iL(𝐳i,λ)\displaystyle={\bm{a}}_{i}^{\mathrm{T}}{\bm{\mu}}_{x,\bar{u}}^{i}-\pazocal{L}({\bm{z}}_{i}^{*},\lambda^{*}) (81a)
=𝒂iT𝝁x,u¯i+(τi)1/2𝚪iT𝐌iT𝒂i𝐒i1.\displaystyle={\bm{a}}_{i}^{\mathrm{T}}{\bm{\mu}}_{x,\bar{u}}^{i}+(\tau^{i})^{1/2}\|\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}{\bm{a}}_{i}\|_{{\bf S}_{i}^{-1}}. (81b)

VIII Robust Non-convex Constraints and Distributed Architecture

VIII-A Initial Approach for robust nonconvex constraint

Let us rewrite the robust nonconvex norm-of-mean constraint (26) as follows

min𝜻iUi𝐀i𝐏ki𝝁x,u¯i+𝐀i𝐏ki𝐌i𝜻i𝒃i2ci,\displaystyle\min_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i}}\|{\bf A}_{i}{\bf P}_{k}^{i}{\bm{\mu}}_{x,\bar{u}}^{i}+{\bf A}_{i}{\bf P}^{i}_{k}{\bf M}_{i}{\bm{\zeta}}^{i}-\bm{b}_{i}\|_{2}\geq c_{i}, (82)

which can be written in terms of 𝜻i{\bm{\zeta}}^{i} as

min𝜻iUi𝜻i𝐅1,kiT(𝐊i)𝜻i+2𝐅2,ki(𝐊i,\displaystyle\min_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i}}{\bm{\zeta}}^{i}{}^{\mathrm{T}}{\bf F}_{1,k}^{i}({\bf K}^{i}){\bm{\zeta}}^{i}+2{\bf F}_{2,k}^{i}({\bf K}^{i}, 𝒖¯i)𝜻i\displaystyle\bar{{\bm{u}}}^{i}){\bm{\zeta}}^{i} (83)
+𝐅3,ki(𝒖¯i)ci2\displaystyle+{\bf F}_{3,k}^{i}(\bar{{\bm{u}}}^{i})\geq c_{i}^{2}

where

𝐅1,ki(𝐊i)\displaystyle{\bf F}_{1,k}^{i}({\bf K}^{i}) =(𝐀i𝐏ki𝐌i)T𝐀i𝐏ki𝐌i,\displaystyle=({\bf A}_{i}{\bf P}^{i}_{k}{\bf M}_{i})^{\mathrm{T}}{\bf A}_{i}{\bf P}^{i}_{k}{\bf M}_{i}, (84a)
𝐅2,ki(𝐊i,𝒖¯i)\displaystyle{\bf F}_{2,k}^{i}({\bf K}^{i},\bar{{\bm{u}}}^{i}) =(𝐀i𝐏ki𝝁x,u¯i𝒃i)T(𝐀i𝐏ki𝐌i),\displaystyle=({\bf A}_{i}{\bf P}_{k}^{i}{\bm{\mu}}_{x,\bar{u}}^{i}-\bm{b}_{i})^{\mathrm{T}}({\bf A}_{i}{\bf P}^{i}_{k}{\bf M}_{i}), (84b)
𝐅3,ki(𝒖¯i)\displaystyle{\bf F}_{3,k}^{i}(\bar{{\bm{u}}}^{i}) =𝐀i𝐏ki𝝁x,u¯i𝒃i22.\displaystyle=\|{\bf A}_{i}{\bf P}_{k}^{i}{\bm{\mu}}_{x,\bar{u}}^{i}-\bm{b}_{i}\|_{2}^{2}. (84c)

The constraint (83) could be reformulated through the following equivalent set of constraints (see [19]),

[𝚪iT𝐅1,ki(𝐊i)𝚪i+α𝐒i𝚪iT𝐅2,ki(𝐊i,u¯i)T𝐅2,ki(𝐊i,u¯i)𝚪i𝐅3,ki(u¯i)ci2ατi]0,\displaystyle\begin{bmatrix}\bm{\Gamma}_{i}^{\mathrm{T}}{\bf F}_{1,k}^{i}({\bf K}^{i})\bm{\Gamma}_{i}+\alpha{\bf S}_{i}&\bm{\Gamma}_{i}^{\mathrm{T}}{\bf F}_{2,k}^{i}({\bf K}^{i},\bar{u}^{i})^{\mathrm{T}}\\ {\bf F}_{2,k}^{i}({\bf K}^{i},\bar{u}^{i})\bm{\Gamma}_{i}&{\bf F}_{3,k}^{i}(\bar{u}^{i})-c_{i}^{2}-\alpha\tau^{i}\end{bmatrix}\succeq 0, (85a)
α0.\displaystyle~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}\alpha\geq 0. (85b)

Nevertheless, the above constraint would still be a nonconvex constraint, as 𝐅1,ki(𝐊i){\bf F}_{1,k}^{i}({\bf K}^{i}), 𝐅2,ki(𝐊i,u¯i){\bf F}_{2,k}^{i}({\bf K}^{i},\bar{u}^{i}), and 𝐅3,ki(u¯i){\bf F}_{3,k}^{i}(\bar{u}^{i}) are nonlinear in 𝐊i{\bf K}^{i}, u¯i\bar{u}^{i}. To address this issue, we first linearize the constraint (82) with respect to 𝐊i,u¯i{\bf K}^{i},\bar{u}^{i} around the nominal 𝐊i,l,u¯i,l{\bf K}^{i,l},\bar{u}^{i,l}, to obtain the following affine robust constraint

𝜻i𝒬T(𝐊i,𝐊i,l)𝜻i+2\displaystyle\bm{\zeta}^{i}{}^{\mathrm{T}}\mathscr{Q}({\bf K}^{i},{\bf K}^{i,l})\bm{\zeta}^{i}+2 𝒬¯(𝐊i,𝐊i,l,𝒖¯i,𝒖¯i,l)𝜻i\displaystyle\bar{\mathscr{Q}}({\bf K}^{i},{\bf K}^{i,l},\bar{{\bm{u}}}^{i},\bar{{\bm{u}}}^{i,l})\bm{\zeta}^{i} (86)
+q(𝒖¯i,𝒖¯i,l)ci2,𝜻iUi\displaystyle+q(\bar{{\bm{u}}}^{i},\bar{{\bm{u}}}^{i,l})\geq c_{i}^{2},\quad\forall\;\bm{\zeta}^{i}\in\pazocal{U}_{i}

where

𝒬=(𝐀i𝐏ki𝐌il)T𝐀i𝐏ki𝐌i\displaystyle\mathscr{Q}=({\bf A}_{i}{\bf P}^{i}_{k}{\bf M}_{i}^{l})^{\mathrm{T}}{\bf A}_{i}{\bf P}^{i}_{k}{\bf M}_{i}
+(𝐀i𝐏ki(𝐌i𝐌il))T𝐀i𝐏ki𝐌il,\displaystyle~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}+({\bf A}_{i}{\bf P}^{i}_{k}({\bf M}_{i}-{\bf M}_{i}^{l}))^{\mathrm{T}}{\bf A}_{i}{\bf P}^{i}_{k}{\bf M}_{i}^{l}, (87a)
𝒬¯=(𝐀i𝐏ki𝝁x,u¯i,l𝒃i)T(𝐀i𝐏ki𝐌i)\displaystyle\bar{\mathscr{Q}}=({\bf A}_{i}{\bf P}_{k}^{i}{\bm{\mu}}_{x,\bar{u}}^{i,l}-\bm{b}_{i})^{\mathrm{T}}({\bf A}_{i}{\bf P}^{i}_{k}{\bf M}_{i})
+(𝐀i𝐏ki(𝝁x,u¯i𝝁x,u¯i,l))T(𝐀i𝐏ki𝐌il),\displaystyle~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}+({\bf A}_{i}{\bf P}_{k}^{i}({\bm{\mu}}_{x,\bar{u}}^{i}-{\bm{\mu}}_{x,\bar{u}}^{i,l}))^{\mathrm{T}}({\bf A}_{i}{\bf P}^{i}_{k}{\bf M}_{i}^{l}), (87b)
q=(𝐀i𝐏ki𝝁x,u¯i,l𝒃i)T(𝐀i𝐏ki(2𝝁x,u¯i𝝁x,u¯i,l)𝒃i).\displaystyle q=({\bf A}_{i}{\bf P}_{k}^{i}{\bm{\mu}}_{x,\bar{u}}^{i,l}-\bm{b}_{i})^{\mathrm{T}}({\bf A}_{i}{\bf P}_{k}^{i}(2{\bm{\mu}}_{x,\bar{u}}^{i}-{\bm{\mu}}_{x,\bar{u}}^{i,l})-\bm{b}_{i}). (87c)

VIII-B Proposition 2 Proof

Let us initially rewrite the constraint (33) as follows

max𝜻iUi𝐀i𝐌~ki𝜻i2c~ici.\max_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i}}\|{\bf A}_{i}\tilde{{\bf M}}_{k}^{i}{\bm{\zeta}}^{i}\|_{2}\leq\tilde{c}_{i}-c_{i}. (88)

The first step is to construct an upper bound for 𝐀i𝐌~ki𝜻i22\|{\bf A}_{i}\tilde{{\bf M}}_{k}^{i}{\bm{\zeta}}^{i}\|_{2}^{2}. Using Proposition 1, we obtain the following two equalities,

max𝜻iUihk,m¯i𝐌iT𝜻i\displaystyle\max_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i}}h_{k,\bar{m}}^{i}{}^{\mathrm{T}}{\bf M}_{i}{\bm{\zeta}}^{i} =(τi)1/2𝚪iT𝐌iThk,m¯i𝐒i1,\displaystyle=(\tau^{i})^{1/2}\|\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}h_{k,\bar{m}}^{i}\|_{{\bf S}_{i}^{-1}}, (89)
min𝜻iUihk,m¯i𝐌iT𝜻i\displaystyle\min_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i}}h_{k,\bar{m}}^{i}{}^{\mathrm{T}}{\bf M}_{i}{\bm{\zeta}}^{i} =(τi)1/2𝚪iT𝐌iThk,m¯i𝐒i1,\displaystyle=-(\tau^{i})^{1/2}\|\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}h_{k,\bar{m}}^{i}\|_{{\bf S}_{i}^{-1}}, (90)

where hk,m¯iTh_{k,\bar{m}}^{i}{}^{\mathrm{T}} is m¯th\bar{m}^{th} row of 𝐀i𝐏ki{\bf A}_{i}{\bf P}_{k}^{i}, for all m¯=1,,m\bar{m}=1,\dots,m. By combining (89) and (90) for every row of 𝐀i𝐌~ki𝜻i{\bf A}_{i}\tilde{{\bf M}}_{k}^{i}{\bm{\zeta}}^{i}, we can arrive to the following relationship,

max𝜻iUi𝐀i𝐌~ki𝜻i2\displaystyle\max_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i}}\|{\bf A}_{i}\tilde{{\bf M}}_{k}^{i}{\bm{\zeta}}^{i}\|_{2} =max𝜻iUi(m¯=1m(hk,m¯i𝐌iT𝜻i)2)1/2\displaystyle=\max_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i}}\bigg{(}\sum_{\bar{m}=1}^{m}(h_{k,\bar{m}}^{i}{}^{\mathrm{T}}{\bf M}_{i}{\bm{\zeta}}^{i})^{2}\bigg{)}^{1/2} (91a)
(m¯=1mmax𝜻iUi(hk,m¯i𝐌iT𝜻i)2)1/2\displaystyle\leq\bigg{(}\sum_{\bar{m}=1}^{m}\max_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i}}(h_{k,\bar{m}}^{i}{}^{\mathrm{T}}{\bf M}_{i}{\bm{\zeta}}^{i})^{2}\bigg{)}^{1/2} (91b)
=(m¯=1mτi𝚪iT𝐌iThk,m¯i𝐒i12)1/2\displaystyle~{}~{}=\bigg{(}\sum_{\bar{m}=1}^{m}\tau^{i}\|\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}h_{k,\bar{m}}^{i}\|_{{\bf S}_{i}^{-1}}^{2}\bigg{)}^{1/2} (91c)

Next, by introducing a variable 𝝁d,ki{\bm{\mu}}_{d,k}^{i} such that

(τi)1/2𝚪iT𝐌iThk,m¯i𝐒i1(𝝁d,ki)m¯,(\tau^{i})^{1/2}\|\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}h_{k,\bar{m}}^{i}\|_{{\bf S}_{i}^{-1}}\leq({\bm{\mu}}_{d,k}^{i})_{\bar{m}}, (92)

for every m¯=1,,m\bar{m}=1,\dots,m, we obtain

max𝜻iUi𝐀i𝐌~ki𝜻i2𝝁d,ki2.\displaystyle\max_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i}}\|{\bf A}_{i}\tilde{{\bf M}}_{k}^{i}{\bm{\zeta}}^{i}\|_{2}\leq\|{\bm{\mu}}_{d,k}^{i}{}\|_{2}. (93)

Therefore, we can consider the following tighter approximation of the constraint (88), given by

𝝁d,ki2c~ici.\displaystyle\|{\bm{\mu}}_{d,k}^{i}{}\|_{2}\leq\tilde{c}_{i}-c_{i}. (94)

VIII-C Proposition 3 Proof

The proof of Proposition 3 is similar to the one of Proposition 2. First, let us rewrite constraint (40) as follows

max𝜻iUi,𝜻jUj𝐀i𝐌~ki𝜻i𝐀j𝐌~kj𝜻j2c~ijcij.\displaystyle\max_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i},~{}{\bm{\zeta}}^{j}\in\pazocal{U}_{j}}\|{\bf A}_{i}\tilde{{\bf M}}_{k}^{i}{\bm{\zeta}}^{i}-{\bf A}_{j}\tilde{{\bf M}}_{k}^{j}{\bm{\zeta}}^{j}\|_{2}\leq\tilde{c}_{ij}-c_{ij}. (95)

The bound on each m¯th\bar{m}^{th} row of 𝐀i𝐌~ki𝜻i𝐀j𝐌~kj𝜻j{\bf A}_{i}\tilde{{\bf M}}_{k}^{i}{\bm{\zeta}}^{i}-{\bf A}_{j}\tilde{{\bf M}}_{k}^{j}{\bm{\zeta}}^{j}, with m¯=1,,m\bar{m}=1,\dots,m, can be given as follows

max𝜻iUi,𝜻jUj(𝐀i𝐌~ki𝜻i𝐀j𝐌~kj𝜻j)m¯=𝝁max,m¯i𝝁min,m¯j\displaystyle\max_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i},~{}{\bm{\zeta}}^{j}\in\pazocal{U}_{j}}({\bf A}_{i}\tilde{{\bf M}}_{k}^{i}{\bm{\zeta}}^{i}-{\bf A}_{j}\tilde{{\bf M}}_{k}^{j}{\bm{\zeta}}^{j})_{\bar{m}}={\bm{\mu}}_{max,\bar{m}}^{i}-{\bm{\mu}}_{min,\bar{m}}^{j} (96a)
min𝜻iUi,𝜻jUj(𝐀i𝐌~ki𝜻i𝐀j𝐌~kj𝜻j)m¯=𝝁min,m¯i𝝁max,m¯j\displaystyle\min_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i},~{}{\bm{\zeta}}^{j}\in\pazocal{U}_{j}}({\bf A}_{i}\tilde{{\bf M}}_{k}^{i}{\bm{\zeta}}^{i}-{\bf A}_{j}\tilde{{\bf M}}_{k}^{j}{\bm{\zeta}}^{j})_{\bar{m}}={\bm{\mu}}_{min,\bar{m}}^{i}-{\bm{\mu}}_{max,\bar{m}}^{j} (96b)

where

𝝁min,m¯i\displaystyle{\bm{\mu}}_{min,\bar{m}}^{i} =min𝜻iUi(𝐀i𝐌~ki𝜻i)m¯,\displaystyle=\min_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i}}({\bf A}_{i}\tilde{{\bf M}}_{k}^{i}{\bm{\zeta}}^{i})_{\bar{m}}, (97)
𝝁max,m¯i\displaystyle{\bm{\mu}}_{max,\bar{m}}^{i} =max𝜻iUi(𝐀i𝐌~ki𝜻i)m¯,\displaystyle=\max_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i}}({\bf A}_{i}\tilde{{\bf M}}_{k}^{i}{\bm{\zeta}}^{i})_{\bar{m}},
𝝁min,m¯j\displaystyle{\bm{\mu}}_{min,\bar{m}}^{j} =min𝜻jUj(𝐀j𝐌~kj𝜻j)m¯,\displaystyle=\min_{{\bm{\zeta}}^{j}\in\pazocal{U}_{j}}({\bf A}_{j}\tilde{{\bf M}}_{k}^{j}{\bm{\zeta}}^{j})_{\bar{m}},
𝝁max,m¯j\displaystyle{\bm{\mu}}_{max,\bar{m}}^{j} =max𝜻jUj(𝐀j𝐌~kj𝜻j)m¯.\displaystyle=\max_{{\bm{\zeta}}^{j}\in\pazocal{U}_{j}}({\bf A}_{j}\tilde{{\bf M}}_{k}^{j}{\bm{\zeta}}^{j})_{\bar{m}}.

Using Proposition 1, we have

𝝁max,m¯i=𝝁min,m¯i=(τi)1/2𝚪iT𝐌iThk,m¯i𝐒i1\displaystyle{\bm{\mu}}_{max,\bar{m}}^{i}=-{\bm{\mu}}_{min,\bar{m}}^{i}=(\tau^{i})^{1/2}\|\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}h_{k,\bar{m}}^{i}\|_{{\bf S}_{i}^{-1}} (98a)
𝝁max,m¯j=𝝁min,m¯j=(τj)1/2𝚪jT𝐌jThk,m¯j𝐒j1\displaystyle{\bm{\mu}}_{max,\bar{m}}^{j}=-{\bm{\mu}}_{min,\bar{m}}^{j}=(\tau^{j})^{1/2}\|\bm{\Gamma}_{j}^{\mathrm{T}}{\bf M}_{j}^{\mathrm{T}}h_{k,\bar{m}}^{j}\|_{{\bf S}_{j}^{-1}} (98b)

where hk,m¯i,Thk,m¯jTh_{k,\bar{m}}^{i}{}^{\mathrm{T}},h_{k,\bar{m}}^{j}{}^{\mathrm{T}} are the m¯th\bar{m}^{th} rows of the matrices 𝐀i𝐏ki,𝐀j𝐏kj{\bf A}_{i}{\bf P}_{k}^{i},{\bf A}_{j}{\bf P}_{k}^{j} respectively. By combining (96) and (98), we obtain

max𝜻iUi,𝜻jUj𝐀i𝐌~ki𝜻i𝐀j𝐌~kj𝜻j2\displaystyle\max_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i},~{}{\bm{\zeta}}^{j}\in\pazocal{U}_{j}}\|{\bf A}_{i}\tilde{{\bf M}}_{k}^{i}{\bm{\zeta}}^{i}-{\bf A}_{j}\tilde{{\bf M}}_{k}^{j}{\bm{\zeta}}^{j}\|_{2} (99)
(m¯=1mmax𝜻iUi,𝜻jUj(𝐀i𝐌~ki𝜻i𝐀j𝐌~kj𝜻j)m¯2)1/2\displaystyle\qquad\leq\bigg{(}\sum_{\bar{m}=1}^{m}\max_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i},~{}{\bm{\zeta}}^{j}\in\pazocal{U}_{j}}({\bf A}_{i}\tilde{{\bf M}}_{k}^{i}{\bm{\zeta}}^{i}-{\bf A}_{j}\tilde{{\bf M}}_{k}^{j}{\bm{\zeta}}^{j})_{\bar{m}}^{2}\bigg{)}^{1/2}
=(m¯=1m(𝝁max,m¯i+𝝁max,m¯j)2)1/2\displaystyle~{}\qquad=\bigg{(}\sum_{\bar{m}=1}^{m}({\bm{\mu}}_{max,\bar{m}}^{i}+{\bm{\mu}}_{max,\bar{m}}^{j})^{2}\bigg{)}^{1/2}

As in the proof of Proposition 2, let us now introduce the variables 𝝁d,ki,𝝁d,kj{\bm{\mu}}_{d,k}^{i},{\bm{\mu}}_{d,k}^{j} such that

(𝝁d,ki)m¯(τi)1/2𝚪iT𝐌iThk,m¯i𝐒i1\displaystyle({\bm{\mu}}_{d,k}^{i})_{\bar{m}}\geq(\tau^{i})^{1/2}\|\bm{\Gamma}_{i}^{\mathrm{T}}{\bf M}_{i}^{\mathrm{T}}h_{k,\bar{m}}^{i}\|_{{\bf S}_{i}^{-1}} (100)
(𝝁d,kj)m¯(τj)1/2𝚪jT𝐌jThk,m¯j𝐒j1\displaystyle({\bm{\mu}}_{d,k}^{j})_{\bar{m}}\geq(\tau^{j})^{1/2}\|\bm{\Gamma}_{j}^{\mathrm{T}}{\bf M}_{j}^{\mathrm{T}}h_{k,\bar{m}}^{j}\|_{{\bf S}_{j}^{-1}} (101)

for every m¯=1,,m\bar{m}=1,\dots,m. Then, we obtain

max𝜻iUi,𝜻jUj𝐀i𝐌~ki𝜻i\displaystyle\max_{{\bm{\zeta}}^{i}\in\pazocal{U}_{i},~{}{\bm{\zeta}}^{j}\in\pazocal{U}_{j}}\|{\bf A}_{i}\tilde{{\bf M}}_{k}^{i}{\bm{\zeta}}^{i} 𝐀j𝐌~kj𝜻j2𝝁d,ki+𝝁d,kj2\displaystyle-{\bf A}_{j}\tilde{{\bf M}}_{k}^{j}{\bm{\zeta}}^{j}\|_{2}\leq\|{\bm{\mu}}_{d,k}^{i}+{\bm{\mu}}_{d,k}^{j}\|_{2} (102)

so we can consider the following tighter approximation to the constraint (95),

𝝁d,ki+𝝁d,kj2c~ijcij.\displaystyle\|{\bm{\mu}}_{d,k}^{i}+{\bm{\mu}}_{d,k}^{j}\|_{2}\leq\tilde{c}_{ij}-c_{ij}. (103)

VIII-D Distributed Architecture Global Updates

The global update steps are given by

{𝝂iu¯,l+1,𝝂id,l+1}i=1N\displaystyle\{{\bm{\nu}}^{\bar{u},l+1}_{i},\;{\bm{\nu}}^{d,l+1}_{i}\}_{i=1}^{N}
=argminLρu¯,ρd({𝝁~iu¯,l+1,𝝁~id,l+1,𝝂~iu¯,𝝂~id,𝝀iu¯,l,𝝀id,l}i=1N)\displaystyle~{}~{}~{}=\operatornamewithlimits{argmin}\;\pazocal{L}_{\rho_{\bar{u}},\rho_{d}}(\{\tilde{{\bm{\mu}}}^{\bar{u},l+1}_{i},\tilde{{\bm{\mu}}}^{d,l+1}_{i},\tilde{{\bm{\nu}}}^{\bar{u}}_{i},\tilde{{\bm{\nu}}}^{d}_{i},{\bm{\lambda}}_{i}^{\bar{u},l},{\bm{\lambda}}_{i}^{d,l}\}_{i=1}^{N})
=argmini=1N𝝀iu¯(𝝁~iu¯𝝂~iu¯)T+𝝀id(𝝁~id𝝂~id)T\displaystyle~{}~{}~{}=\operatornamewithlimits{argmin}\;\sum_{i=1}^{N}{\bm{\lambda}}_{i}^{\bar{u}}{}^{\mathrm{T}}(\tilde{{\bm{\mu}}}^{\bar{u}}_{i}-\tilde{{\bm{\nu}}}^{\bar{u}}_{i})+{\bm{\lambda}}_{i}^{d}{}^{\mathrm{T}}(\tilde{{\bm{\mu}}}^{d}_{i}-\tilde{{\bm{\nu}}}^{d}_{i}) (104)
+ρu¯2𝝁~iu¯𝝂~iu¯22+ρd2𝝁~id𝝂~id22\displaystyle~{}~{}~{}~{}\qquad\qquad+\frac{\rho_{\bar{u}}}{2}\|\tilde{{\bm{\mu}}}^{\bar{u}}_{i}-\tilde{{\bm{\nu}}}^{\bar{u}}_{i}\|_{2}^{2}+\frac{\rho_{d}}{2}\|\tilde{{\bm{\mu}}}^{d}_{i}-\tilde{{\bm{\nu}}}^{d}_{i}\|_{2}^{2}

This update can be decoupled for each agent ii as follows

𝝂iu¯,l+1,𝝂id,l+1\displaystyle{\bm{\nu}}^{\bar{u},l+1}_{i},{\bm{\nu}}^{d,l+1}_{i} (105)
=argmin𝝂iu¯,𝝂id𝝀¯iu¯,l(𝝁^iu¯,l+1𝝂iu¯)T+𝝀¯id,l(𝝁id,l+1𝝂id)T\displaystyle~{}~{}~{}=\operatornamewithlimits{argmin}_{{\bm{\nu}}^{\bar{u}}_{i},{\bm{\nu}}^{d}_{i}}\;\bar{{\bm{\lambda}}}_{i}^{\bar{u},l}{}^{\mathrm{T}}(\hat{{\bm{\mu}}}^{\bar{u},l+1}_{i}-{\bm{\nu}}^{\bar{u}}_{i})+\bar{{\bm{\lambda}}}_{i}^{d,l}{}^{\mathrm{T}}({\bm{\mu}}^{d,l+1}_{i}-{\bm{\nu}}^{d}_{i})
+ρu¯2𝝁^iu¯,l+1𝝂iu¯22+ρd2𝝁id,l+1𝝂id22\displaystyle~{}~{}~{}\qquad\qquad+\frac{\rho_{\bar{u}}}{2}\|\hat{{\bm{\mu}}}^{\bar{u},l+1}_{i}-{\bm{\nu}}^{\bar{u}}_{i}\|_{2}^{2}+\frac{\rho_{d}}{2}\|{\bm{\mu}}^{d,l+1}_{i}-{\bm{\nu}}^{d}_{i}\|_{2}^{2}
+j^Pi(𝝀¯j^i,u¯,l(𝝁¯j^i,u¯,l+1𝝂iu¯)T\displaystyle~{}~{}~{}\qquad\qquad+\sum_{\hat{j}\in\pazocal{P}_{i}}\big{(}\bar{{\bm{\lambda}}}_{\hat{j}}^{i,\bar{u},l}{}^{\mathrm{T}}(\bar{{\bm{\mu}}}^{i,\bar{u},l+1}_{\hat{j}}-{\bm{\nu}}^{\bar{u}}_{i})
+𝝀¯j^i,d,l(𝝁¯j^i,d,l+1𝝂id)T+ρu¯2𝝁¯j^i,u¯,l+1𝝂iu¯22\displaystyle~{}~{}~{}\qquad\qquad+\bar{{\bm{\lambda}}}_{\hat{j}}^{i,d,l}{}^{\mathrm{T}}(\bar{{\bm{\mu}}}^{i,d,l+1}_{\hat{j}}-{\bm{\nu}}^{d}_{i})+\frac{\rho_{\bar{u}}}{2}\|\bar{{\bm{\mu}}}^{i,\bar{u},l+1}_{\hat{j}}-{\bm{\nu}}^{\bar{u}}_{i}\|_{2}^{2}
+ρd2𝝁¯j^i,d,l+1𝝂id22)\displaystyle~{}~{}~{}\qquad\qquad+\frac{\rho_{d}}{2}\|\bar{{\bm{\mu}}}^{i,d,l+1}_{\hat{j}}-{\bm{\nu}}^{d}_{i}\|_{2}^{2}\big{)}

To find the minimum on the RHS, let us differentiate the RHS with respect to the global variables 𝝂iu¯,𝝂id{\bm{\nu}}^{\bar{u}}_{i},{\bm{\nu}}^{d}_{i} and equate it to zero. For 𝝂iu¯,l+1{\bm{\nu}}^{\bar{u},l+1}_{i}, we get

𝝀¯iu¯,l+ρu¯(𝝁^iu¯,l+1𝝂iu¯)\displaystyle-\bar{{\bm{\lambda}}}_{i}^{\bar{u},l}+\rho_{\bar{u}}(\hat{{\bm{\mu}}}^{\bar{u},l+1}_{i}-{\bm{\nu}}^{\bar{u}}_{i}) (106)
+j^Pi(𝝀¯j^i,u¯,l+ρu¯(𝝁¯j^i,u¯,l+1𝝂iu¯))=0\displaystyle\qquad+\sum_{\hat{j}\in\pazocal{P}_{i}}\big{(}-\bar{{\bm{\lambda}}}_{\hat{j}}^{i,\bar{u},l}+\rho_{\bar{u}}(\bar{{\bm{\mu}}}^{i,\bar{u},l+1}_{\hat{j}}-{\bm{\nu}}^{\bar{u}}_{i})\big{)}=0

which leads to

𝝂iu¯,l+1\displaystyle{\bm{\nu}}^{\bar{u},l+1}_{i} =1(n(Pi)+1)(𝝀¯iu¯,lρu¯+𝝁^iu¯,l+1\displaystyle=\frac{1}{(n(\pazocal{P}_{i})+1)}\bigg{(}\frac{\bar{{\bm{\lambda}}}_{i}^{\bar{u},l}}{\rho_{\bar{u}}}+\hat{{\bm{\mu}}}_{i}^{\bar{u},l+1} (107)
+j^Pi(𝝀¯j^i,u¯,lρu¯+𝝁¯j^i,u¯,l+1))\displaystyle\qquad\qquad+\sum_{\hat{j}\in\pazocal{P}_{i}}\big{(}\frac{\bar{{\bm{\lambda}}}_{\hat{j}}^{i,\bar{u},l}}{\rho_{\bar{u}}}+\bar{{\bm{\mu}}}_{\hat{j}}^{i,\bar{u},l+1}\big{)}\bigg{)}

The update for 𝝂id,l+1{\bm{\nu}}^{d,l+1}_{i} can be derived similarly.

IX Simulation Data

IX-A System Parameters

Each agent iVi\in\pazocal{V} is subject to the following 2D double integrator dynamics,

Aki=[100.0500100.0500100001],Bki=[0.0013000.00130.05000.05],\displaystyle A_{k}^{i}=\begin{bmatrix}1&0&0.05&0\\ 0&1&0&0.05\\ 0&0&1&0\\ 0&0&0&1\end{bmatrix},\quad B_{k}^{i}=\begin{bmatrix}0.0013&0\\ 0&0.0013\\ 0.05&0\\ 0&0.05\end{bmatrix},
Cki=randn(4,2),Dki=𝐈4×4\displaystyle C_{k}^{i}=\texttt{randn}(4,2),\quad D_{k}^{i}={\bf I}_{4\times 4}

for k0,T1k\in\llbracket 0,T-1\rrbracket. Further, we enforce that CkiF=1\|C_{k}^{i}\|_{F}=1. We consider 𝐒i=𝐈,𝐏i=𝐈{\bf S}_{i}={\bf I},{\bf P}_{i}={\bf I} for all agents. In each experiment, we use τi\tau^{i} as the uncertainty level. The cost matrices are set to be 𝐑u¯=0.05𝐈{\bf R}_{\bar{u}}=0.05{\bf I}, and 𝐑𝐊=0.05𝐈{\bf R}_{{\bf K}}=0.05{\bf I}.

IX-B ADMM Parameters

The penalty parameters are set as ρu¯=100\rho_{\bar{u}}=100, ρd=1\rho_{d}=1, with maximum ADMM iterations set to be 200. Further, we consider the following primal and dual residual conditions as the convergence criteria for ADMM at the lthl^{th} iteration. The primal residual condition is given by

(i=1N𝝁~iu¯,l𝝂~iu¯,l2+𝝁~id,l𝝂~id,l2N+i=1Nn(Ni))1/2ϵprimal\bigg{(}\frac{\sum_{i=1}^{N}\|\tilde{{\bm{\mu}}}^{\bar{u},l}_{i}-\tilde{{\bm{\nu}}}^{\bar{u},l}_{i}\|^{2}+\|\tilde{{\bm{\mu}}}^{d,l}_{i}-\tilde{{\bm{\nu}}}^{d,l}_{i}\|^{2}}{N+\sum_{i=1}^{N}n(\pazocal{N}_{i})}\bigg{)}^{1/2}\leq\epsilon_{\text{primal}} (108)

while the dual residual one is

(i=1Nρu¯2𝝂iu¯,l𝝂iu¯,l12+ρd2𝝂id,l𝝂id,l12)1/2Nϵdual\frac{\big{(}\sum_{i=1}^{N}\rho_{\bar{u}}^{2}\|{\bm{\nu}}^{\bar{u},l}_{i}-{\bm{\nu}}^{\bar{u},l-1}_{i}\|^{2}+\rho_{d}^{2}\|{\bm{\nu}}^{d,l}_{i}-{\bm{\nu}}^{d,l-1}_{i}\|^{2}\big{)}^{1/2}}{N}\leq\epsilon_{\text{dual}}

Finally, we set both residual thresholds ϵprimal,ϵdual=1.0\epsilon_{\text{primal}},\epsilon_{\text{dual}}=1.0.

IX-C Constraint Parameters

For both inter-agent constraints and obstacle avoidance constraints, we use 𝐇pos=[𝐈2×20]{\bf H}_{pos}=\begin{bmatrix}{\bf I}_{2\times 2}&0\end{bmatrix}. The collision distance threshold for inter-agent and obstacle avoidance constraints are set at 0.25, 0.5 respectively, for the experiments (1, 3, 4). For experiment (2), both inter-agent and obstacle collision distance threshold are set at 0.25. Further, in the experiments (3, 4) involving chance constraints, the probability threshold of chance constraints is set to be p=0.05p=0.05.